batbot: a biologically inspired flapping and morphing bat robot

223
BaTboT: a biologically inspired flapping and morphing bat robot actuated by SMA-based artificial muscles. Julian David Colorado M. Department of Electronics, Informatics and Industrial Engineering Universidad Politecnica de Madrid, Spain A thesis submitted for the degree of Doctor of Philosophy in Robotics 2012

Upload: vonga

Post on 11-Feb-2017

229 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: BaTboT: a biologically inspired flapping and morphing bat robot

BaTboT: a biologically inspired flapping and

morphing bat robot actuated by SMA-based

artificial muscles.

Julian David Colorado M.

Department of Electronics, Informatics and Industrial Engineering

Universidad Politecnica de Madrid, Spain

A thesis submitted for the degree of

Doctor of Philosophy in Robotics

2012

Page 2: BaTboT: a biologically inspired flapping and morphing bat robot

Title:

BaTboT: a biologically inspired flapping and morphing bat robot actuated by

SMA-based artificial muscles.

Author:

Julian David Colorado M, M.Sc

Director:

Prof. Antonio Barrientos Cruz, Ph.D

Prof. Claudio Rossi, Ph.D

Robotics and Cybernetics Group

Tribunal nombrado por el Mgfco. y Excmo. Sr. Rector de la Universidad Politectica de

Madrid, el dıa ........ de ........ de 2012.

Tribunal

Presidente:

Vocal:

Vocal:

Vocal:

Secretario:

Suplente:

Suplente:

Realizado el acto de lectura y defensa de la Tesis el dıa ........ de .......de 2012.

Calificacion de la Tesis: .......

El presidente: Los Vocales:

El Secretario:

ii

Page 3: BaTboT: a biologically inspired flapping and morphing bat robot

Abstract

Bats are animals that posses high maneuvering capabilities. Their wings contain

dozens of articulations that allow the animal to perform aggressive maneuvers by

means of controlling the wing shape during flight (morphing-wings). There is no

other flying creature in nature with this level of wing dexterity and there is biological

evidence that the inertial forces produced by the wings have a key role in the attitude

movements of the animal.

This can inspire the design of highly articulated morphing-wing micro

air vehicles (not necessarily bat-like) with a significant wing-to-body

mass ratio.

This thesis presents the development of a novel bat-like micro air vehicle (BaTboT )

inspired by the morphing-wing mechanism of bats. BaTboT’s morphology is alike

in proportion compared to its biological counterpart Cynopterus brachyotis, which

provides the biological foundations for developing accurate mathematical models

and methods that allow for mimicking bat flight.

In nature bats can achieve an amazing level of maneuverability by combining flap-

ping and morphing wingstrokes. Attempting to reproduce the biological wing ac-

tuation system that provides that kind of motion using an artificial counterpart

requires the analysis of alternative actuation technologies more likely muscle fiber

arrays instead of standard servomotor actuators. Thus, NiTinol Shape Memory Al-

loys (SMAs) acting as artificial biceps and triceps muscles are used for mimicking

the morphing wing mechanism of the bat flight apparatus. This antagonistic con-

figuration of SMA-muscles response to an electrical heating power signal to operate.

This heating power is regulated by a proper controller that allows for accurate and

fast SMA actuation. Morphing-wings will enable to change wings geometry with

the unique purpose of enhancing aerodynamics performance. During the downstroke

Page 4: BaTboT: a biologically inspired flapping and morphing bat robot

phase of the wingbeat motion both wings are fully extended aimed at increasing the

area surface to properly generate lift forces. Contrary during the upstroke phase

of the wingbeat motion both wings are retracted to minimize the area and thus

reducing drag forces.

Morphing-wings do not only improve on aerodynamics but also on the inertial forces

that are key to maneuver. Thus, a modeling framework is introduced for analyzing

how BaTboT should maneuver by means of changing wing morphology. This allows

the definition of requirements for achieving forward and turning flight according

to the kinematics of the wing modulation. Motivated by the biological fact about

the influence of wing inertia on the production of body accelerations, an attitude

controller is proposed. The attitude control law incorporates wing inertia informa-

tion to produce desired roll (φ) and pitch (θ) acceleration commands. This novel

flight control approach is aimed at incrementing net body forces (Fnet) that generate

propulsion.

Mimicking the way how bats take advantage of inertial and

aerodynamical forces produced by the wings in order to both increase

lift and maneuver is a promising way to design more efficient

flapping/morphing wings MAVs. The novel wing modulation strategy

and attitude control methodology proposed in this thesis provide a

totally new way of controlling flying robots, that eliminates the need of

appendices such as flaps and rudders, and would allow performing

more efficient maneuvers, especially useful in confined spaces.

As a whole, the BaTboT project consists of five major stages of development:

• Study and analysis of biological bat flight data reported in specialized

literature aimed at defining design and control criteria.

• Formulation of mathematical models for: i) wing kinematics, ii) dynamics,

iii) aerodynamics, and iv) SMA muscle-like actuation. It is aimed at modeling

the effects of modulating wing inertia into the production of net body forces

for maneuvering.

Page 5: BaTboT: a biologically inspired flapping and morphing bat robot

• Bio-inspired design and fabrication of: i) skeletal structure of wings and

body, ii) SMA muscle-like mechanisms, iii) the wing-membrane, and iv) elec-

tronics onboard. It is aimed at developing the bat-like platform (BaTboT)

that allows for testing the methods proposed.

• The flight controller: i) control of SMA-muscles (morphing-wing modula-

tion) and ii) flight control (attitude regulation). It is aimed at formulating

the proper control methods that allow for the proper modulation of BaTboT’s

wings.

• Experiments: it is aimed at quantifying the effects of properly wing modu-

lation into aerodynamics and inertial production for maneuvering. It is also

aimed at demonstrating and validating the hypothesis of improving flight effi-

ciency thanks to the novel control methods presented in this thesis.

This thesis introduces the challenges and methods to address these stages. Wind-

tunnel experiments will be oriented to discuss and demonstrate how the wings can

considerably affect the dynamics/aerodynamics of flight and how to take advan-

tage of wing inertia modulation that the morphing-wings enable to properly change

wings’ geometry during flapping.

Page 6: BaTboT: a biologically inspired flapping and morphing bat robot

Resumen:

Los murcielagos son mamıferos con una alta capacidad de maniobra. Sus alas estan

conformadas por docenas de articulaciones que permiten al animal maniobrar gracias

al cambio geometrico de las alas durante el vuelo. Esta caracterıstica es conocida

como (alas morficas). En la naturaleza, no existe ningun especimen volador con se-

mejante grado de dexteridad de vuelo, y se ha demostrado, que las fuerzas inerciales

producidas por el batir de las alas juega un papel fundamental en los movimientos

que orientan al animal en vuelo.

Estas caracterısticas pueden inspirar el diseno de un micro vehıculo

aereo compuesto por alas morficas con redundantes grados de libertad,

y cuya proporcion entre la masa de sus alas y el cuerpo del robot sea

significativa.

Esta tesis doctoral presenta el desarrollo de un novedoso robot aereo inspirado en

el mecanismo de ala morfica de los murcielagos. El robot, llamado BaTboT, ha sido

disenado con parametros morfologicos muy similares a los descritos por su sımil

biologico Cynopterus brachyotis. El estudio biologico de este especimen ha permi-

tido la definicion de criterios de diseno y modelos matematicos que representan el

comportamiento del robot, con el objetivo de imitar lo mejor posible la biomecanica

de vuelo de los murcielagos.

La biomecanica de vuelo esta definida por dos tipos de movimiento de las alas: ale-

teo y cambio de forma. Intentar imitar como los murcielagos cambian la forma de

sus alas con un prototipo artificial, requiere el analisis de metodos alternativos de

actuacion que se asemejen a la biomecanica de los musculos que actuan las alas, y

evitar el uso de sistemas convencionales de actuacion como servomotores o motores

DC. En este sentido, las aleaciones con memoria de forma, o por sus siglas en ingles

(SMA), las cuales son fibras de NiTinol que se contraen y expanden ante estımulos

termicos, han sido usados en este proyecto como musculos artificiales que actuan

como bıceps y trıceps de las alas, proporcionando la funcionalidad de ala morfica

previamente descrita. De esta manera, los musculos de SMA son mecanicamente

Page 7: BaTboT: a biologically inspired flapping and morphing bat robot

posicionados en una configuracion antagonista que permite la rotacion de las artic-

ulaciones del robot. Los actuadores son accionados mediante una senal de potencia

la cual es regulada por un sistema de control encargado que los musculos de SMA

respondan con la precision y velocidad deseada. Este sistema de control morfico de

las alas permitira al robot cambiar la forma de las mismas con el unico proposito de

mejorar el desempeno aerodinamico. Durante la fase de bajada del aleteo, las alas

deben estar extendidas para incrementar la produccion de fuerzas de sustentacion.

Al contrario, durante el ciclo de subida del aleteo, las alas deben contraerse para

minimizar el area y reducir las fuerzas de friccion aerodinamica.

El control de alas morficas no solo mejora el desempeno aerodinamico, tambien

impacta la generacion de fuerzas inerciales las cuales son esenciales para maniobrar

durante el vuelo. Con el objetivo de analizar como el cambio de geometrıa de las alas

influye en la definicion de maniobras y su efecto en la produccion de fuerzas netas,

simulaciones y experimentos han sido llevados a cabo para medir como distintos

patrones de modulacion de las alas influyen en la produccion de aceleraciones lineales

y angulares. Gracias a estas mediciones, se propone un control de vuelo, o control

de actitud, el cual incorpora informacion inercial de las alas para la definicion de

referencias de aceleracion angular. El objetivo de esta novedosa estrategia de control

radica en el incremento de fuerzas netas para la adecuada generacion de movimiento

(Fnet).

Imitar como los murcielagos ajustan sus alas con el proposito de

incrementar las fuerzas de sustentacion y mejorar la maniobra en

vuelo es definitivamente un topico de mucho interes para el diseno de

robots aeros mas eficientes. La propuesta de control de vuelo definida

en este trabajo de investigacion podrıa dar paso a una nueva forma de

control de vuelo de robots aereos que no necesitan del uso de partes

mecanicas tales como alerones, etc. Este control tambien permitirıa el

desarrollo de vehıculos con mayor capacidad de maniobra.

El desarrollo de esta investigacion se centra en cinco etapas:

• Estudiar y analizar el vuelo de los murcielagos con el proposito de definir

criterios de diseno y control.

Page 8: BaTboT: a biologically inspired flapping and morphing bat robot

• Formular modelos matematicos que describan la: i) cinematica de las alas,

ii) dinamica, iii) aerodinamica, y iv) actuacion usando SMA. Estos modelos

permiten estimar la influencia de modular las alas en la produccion de fuerzas

netas.

• Diseno y fabricacion de BaTboT: i) estructura de las alas y el cuerpo, ii)

mecanismo de actuacion morfico basado en SMA, iii) membrana de las alas, y

iv) electronica abordo.

• Contro de vuelo compuesto por: i) control de la SMA (modulacion de las

alas) y ii) regulacion de maniobra (actitud).

• Experimentos: estan enfocados en poder cuantificar cuales son los efectos

que ejercen distintos perfiles de modulacion del ala en el comportamiento

aerodinamico e inercial. El objetivo es demostrar y validar la hipotesis planteada

al inicio de esta investigacion: mejorar eficiencia de vuelo gracias al novedoso

control de orientacion (actitud) propuesto en este trabajo.

A lo largo del desarrollo de cada una de las cinco etapas, se iran presentando los retos,

problematicas y soluciones a abordar. Los experimentos son realizados utilizando

un tunel de viento con la instrumentacion necesaria para llevar a cabo las mediciones

de desempeno respectivas. En los resultados se discutira y demostrara que la inercia

producida por las alas juega un papel considerable en el comportamiento dinamico

y aerodinamico del sistema y como poder tomar ventaja de dicha caracterıstica para

regular patrones de modulacion de las alas que conduzcan a mejorar la eficiencia

del robot en futuros vuelos.

Page 9: BaTboT: a biologically inspired flapping and morphing bat robot

To God, my beloved father, his blessing is the reason for who I am nowadays. To

my parents, German and Patricia, who have always loved and supported me

during my whole life, their unconditional love, patience, wisdom and advising have

made me go beyond my aims. To my beloved brother Juan Felipe and Marita

because of their unconditional love and tenderness throughout my life. To my

advisors and friends, Antonio Barrientos and Claudio Rossi, who have offered me

constructive advice and helped me focus on feasible subjects to deal within this

endeavor of making this research successful. Finally, to all friends and fellows,

thank you very much, and The Lord bless you....

Page 10: BaTboT: a biologically inspired flapping and morphing bat robot

Acknowledgements

Words are not enough to express my gratitude and to acknowledge to whom anyhow

contribute during the development of this endeavor research. The author would like

to thank to professors Kenny Breuer and Sharon Swartz for providing the support

and useful knowledge about the robot design, bat flight kinematics and aerody-

namics. To the Breuer Lab team for providing the wind-tunnel facility of Brown

University, USA and their support with the experiments. To the Robotics and

Cybernetics Group team for their warm collaborative environment and helpful dis-

cussions in general robotic sciences. Last but not least, to the Community of Madrid

and Universidad Politecnica de Madrid for their funding during the development of

this research.

Page 11: BaTboT: a biologically inspired flapping and morphing bat robot

Contents

List of Figures viii

List of Tables xxi

1 Introduction 1

1.1 The problem and motivations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.2 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

1.3 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

1.4 Original Contributions of this Work . . . . . . . . . . . . . . . . . . . . . . . . . 10

1.5 Thesis outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2 Literature Review 13

2.1 General Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2.2 Nature flyers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.2.1 Biomechanics: insects, birds, and bats . . . . . . . . . . . . . . . . . . . . 14

2.2.2 Bat biology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.3 Bat flight research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

2.4 Shape Memory Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2.4.1 Basic foundations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

2.4.2 Advantages and drawbacks . . . . . . . . . . . . . . . . . . . . . . . . . . 30

2.5 Morphing-wing MAVs with smart actuation . . . . . . . . . . . . . . . . . . . . . 32

2.6 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

3 From bats to BaTboT: Mimicking biology 39

3.1 General overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

3.2 Review on biological flight performance data . . . . . . . . . . . . . . . . . . . . 39

3.2.1 Measurements of wing morphological parameters . . . . . . . . . . . . . . 40

iii

Page 12: BaTboT: a biologically inspired flapping and morphing bat robot

CONTENTS

3.2.2 Measurements of kinematics parameters . . . . . . . . . . . . . . . . . . . 41

3.2.3 Measurements of Aerodynamics parameters . . . . . . . . . . . . . . . . . 42

3.3 Choice of species . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

3.3.1 Wing morphology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

3.3.2 Biological-based framework for modeling and design . . . . . . . . . . . . 46

3.4 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

4 BaTboT modeling 48

4.1 General overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

4.2 Kinematics model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

4.2.1 Topology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

4.2.2 Wing and body kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . 50

4.2.3 Wing trajectories and manuevers . . . . . . . . . . . . . . . . . . . . . . . 54

4.3 Inertial model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

4.3.1 Spatial notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

4.3.2 Equations of motion (EoM) . . . . . . . . . . . . . . . . . . . . . . . . . . 63

4.3.3 Rolling and pitching torques . . . . . . . . . . . . . . . . . . . . . . . . . 65

4.4 Aerodynamics model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

4.4.1 Lift and drag forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

4.4.2 Net forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

4.5 SMA muscle-like wing actuation . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

4.5.1 SMA actuation configurations . . . . . . . . . . . . . . . . . . . . . . . . . 70

4.5.2 SMA phenomenological model . . . . . . . . . . . . . . . . . . . . . . . . 70

4.6 Simulation and experimental results . . . . . . . . . . . . . . . . . . . . . . . . . 74

4.6.1 Open-loop simulator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

4.6.2 Wing torques for actuation . . . . . . . . . . . . . . . . . . . . . . . . . . 75

4.6.3 Body torques for maneuvering: experiments for inertial model validation . 82

4.6.4 SMA-muscle limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

4.7 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

5 BaTboT design and Fabrication 90

5.1 The general method for BaTboT’s design . . . . . . . . . . . . . . . . . . . . . . 90

5.2 Prototype characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

5.2.1 Design process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

iv

Page 13: BaTboT: a biologically inspired flapping and morphing bat robot

CONTENTS

5.2.2 Components and weight distribution . . . . . . . . . . . . . . . . . . . . . 92

5.3 BaTboT mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

5.3.1 Step 2: Design criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

5.3.1.1 Bio-inspired parameters . . . . . . . . . . . . . . . . . . . . . . . 96

5.3.1.2 Actuators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

5.3.2 Step 3: Fixed-wing design . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

5.3.2.1 Flapping-wing mechanism . . . . . . . . . . . . . . . . . . . . . . 101

5.3.2.2 Membrane issues . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

5.3.3 Step 4: Articulated-wing design . . . . . . . . . . . . . . . . . . . . . . . . 103

5.3.4 Step 5: Morphing-wing mechanism . . . . . . . . . . . . . . . . . . . . . . 105

5.3.5 Step 6: The wing-membrane . . . . . . . . . . . . . . . . . . . . . . . . . 107

5.4 BaTboT electronics and sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

5.4.1 Arduino controller-board . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

5.4.2 The Inertial Measurement Unit (IMU) . . . . . . . . . . . . . . . . . . . . 113

5.4.3 SMA power drivers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

5.5 BaTboT consumption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

5.6 BaTboT costs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

5.7 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

6 BaTboT Control 118

6.1 Control goal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

6.2 Flight Control Architecture (FCA) . . . . . . . . . . . . . . . . . . . . . . . . . . 119

6.3 SMA actuation: experimental characterization . . . . . . . . . . . . . . . . . . . 122

6.3.1 Frequency analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

6.3.2 Experimental validation of SMA actuation model . . . . . . . . . . . . . . 126

6.3.3 Data summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

6.4 SMA Resistance-to-Motion relationship (RM) . . . . . . . . . . . . . . . . . . . . 131

6.5 Morphing-wing control (inner loop) . . . . . . . . . . . . . . . . . . . . . . . . . . 132

6.5.1 Sliding-mode control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

6.5.2 PID control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

6.5.3 Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

6.5.4 Morphing-wing control algorithm . . . . . . . . . . . . . . . . . . . . . . . 136

6.6 Attitude control (outer loop) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

6.6.1 Backstepping+DAF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

v

Page 14: BaTboT: a biologically inspired flapping and morphing bat robot

CONTENTS

6.6.2 Attitude control algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

6.7 Simulation and experimental results . . . . . . . . . . . . . . . . . . . . . . . . . 143

6.7.1 Variables and parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

6.7.2 Morphing-wing control response . . . . . . . . . . . . . . . . . . . . . . . 145

6.7.3 Attitude control response . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

6.8 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

7 General experimental results 154

7.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

7.1.1 Methods and goals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

7.1.2 The wind-tunnel setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

7.2 Control performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

7.2.1 Morphing-wing control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

7.2.1.1 SMA accuracy and speed . . . . . . . . . . . . . . . . . . . . . . 158

7.2.1.2 SMA fatigue issues . . . . . . . . . . . . . . . . . . . . . . . . . . 159

7.2.2 Attitude control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

7.2.2.1 Forward flight . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

7.2.2.2 Turning flight . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

7.3 Aerodynamics experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163

7.4 Efficient Flight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

7.5 Discussion of results: Towards efficient flight . . . . . . . . . . . . . . . . . . . . . 168

7.5.1 Morphing-wing modulation . . . . . . . . . . . . . . . . . . . . . . . . . . 168

7.5.2 Wing inertia for efficient flight . . . . . . . . . . . . . . . . . . . . . . . . 168

8 Conclusions and Future Work 170

8.1 General conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170

8.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172

8.3 Thesis schedule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173

9 Publications 174

9.1 Journals, book chapters and conference proceedings . . . . . . . . . . . . . . . . 174

9.2 Press and media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175

References 176

vi

Page 15: BaTboT: a biologically inspired flapping and morphing bat robot

CONTENTS

10 Annexes 180

10.1 Floating-base forward kinematics Matlab-code . . . . . . . . . . . . . . . . . . . . 180

10.2 Floating-base inverse dynamics Matlab-code . . . . . . . . . . . . . . . . . . . . . 181

10.3 Floating-base forward dynamics Matlab-code . . . . . . . . . . . . . . . . . . . . 181

10.4 SMA phenomenological model Matlab-code . . . . . . . . . . . . . . . . . . . . . 182

10.5 Control code programmed into the Arduino . . . . . . . . . . . . . . . . . . . . . 184

10.6 Flight Control Matlab environment . . . . . . . . . . . . . . . . . . . . . . . . . . 189

vii

Page 16: BaTboT: a biologically inspired flapping and morphing bat robot

List of Figures

1.1 BaTboT. The overall mass of the skeleton+electronics+battery is 125g. The wingspan:

53cm (wings fully extended). Each wing of the robot has six degrees of freedom (dof):

2-dof at shoulder, 1-dof at elbow, and 3-dof at wrist joint. The body frame {b} is a

6-dof floating body. Rotations about the body-frame {b}-xb, yb, zb axes are designated

roll, pitch and yaw following aerodynamic conventions. Frame {o} is the inertial frame. 2

1.2 a) bats are agile flyer for hunting preys in air, water or even ground, b) bats

can hover like hummingbirds, consuming less energy (10 − 20Hz of flapping

frequency), c) bat wings can camber, stretch, extend and fold like no other

flying animal in nature, d) VTOL flight capacity, Source: Breuer Lab, http:

//brown.edu/Research/Breuer-Lab/research/batflight.html . . . . . . . . 5

1.3 Structural steps to be followed during the thesis aimed at the development of

BaTboT. The pictures depicted herein, correspond to the final BaTboT proto-

type. The forthcoming chapters will introduce each step with all the details.

Source: The author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.1 Structure of animal wings showing the main skeletal support. (Left) vertebrates,

(Right) insects. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.2 Comparison for several species of small-scale flyers: a) Wingspan, b) Aspect

ratio, c) Wing loading, d) Wingbeat frequency. Source from (1), (2), (3). . . . . . 16

2.3 a) Cartoon of aerodynamics forces acting on a typical wing design, b) Lift and

drag changes with the angle of attack for a typical wing design. c) Comparison

of the lift to drag ratio. d) Power Flight. Source from (1), (2), (3). . . . . . . . . 18

2.4 Bat anatomy. A) body anatomy, B) Pectoral skeleton, C) Dorsal view of re-

tracted right arm, D) Dorsal view of expanded right arm. Source (4). . . . . . . . 20

2.5 Bat muscle structure to power wingstroke motion. . . . . . . . . . . . . . . . . . 21

viii

Page 17: BaTboT: a biologically inspired flapping and morphing bat robot

LIST OF FIGURES

2.6 Bat wing membrane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2.7 Tracking results of a bat using a 52 degree of freedom articulated model. Shown

on top are frames extracted from high speed video of a landing bat. Shown on

the bottom are the corresponding frames of the reconstructed three-dimensional

wing and body kinematics. –Caption extracted from (5)–. . . . . . . . . . . . . . 23

2.8 Reconstruction of the wake of C. brachyotis bat. Source: (6). . . . . . . . . . . . 24

2.9 Strain measurements of bat’s wing membrane. Source: (7). . . . . . . . . . . . . 25

2.10 Sensory wing hair before (A) and after (B and C) depilation. (A) Scanning

electron microscope image from a domed hair located on the ventral trailing

edge (location is marked by a gray circle in schematic to the Right) of Eptesicus

fuscus. Caption extracted from Source: (8). . . . . . . . . . . . . . . . . . . . . . 25

2.11 a) The hysteresis curve of SMA, b) stress-strain-temperature curve of SMA ex-

hibiting the one-way shape memory effect, c) stress-strain-temperature curve of

SMA exhibiting the two-way shape memory effect, d) stress-strain-temperature

curve of SMA exhibiting pseudoelasticity behavior. Source: (9). . . . . . . . . . . 27

2.12 a) The BATMAV robot, b) Detailed arm assembly using SMAs-bases muscles.

Source: (10). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

2.13 a) Bat-wing platform in the wind-tunnel, b) Nylon membrane, c) Spandex mem-

brane, d) Silicone membrane. Source: (11). . . . . . . . . . . . . . . . . . . . . . 33

2.14 Harvard RoboBee, actuated by PZT. Source: (12). . . . . . . . . . . . . . . . . . 34

2.15 Inner structure and possible control modes of a multi-functional trailing edge.

Source: (13). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

2.16 Layout of an MFC actuation device for MAV morphing-wing camber. Source:

(14). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

2.17 TOP (left to right) Smart-bird by FESTO, inspired by the Seagles (15), Combat

by University of Michigan, inspired by bat’s navigation system (16), DARPA

nano-air hummingbird (http://www.darpa.mil). BELOW (left to right) The

Gull Wing by (17), MFX-1 by NextGen Aeronautics, inspired by the batwing

internal structure (18), Prototype of Entomoter MAV (19). . . . . . . . . . . . . 36

ix

Page 18: BaTboT: a biologically inspired flapping and morphing bat robot

LIST OF FIGURES

3.1 Right lateral view of the bat with respect to the inertial coordinate system {o}.Green dots are the path of the wrist joint whereas red dots are the path of the

wingtip over a wingbeat cycle. The position and posture of the right wing are

shown at three time points in the wingbeat cycle. Source: (20). . . . . . . . . . 40

3.2 (A) Maximum wingspan, (B) minimum wingspan, (C) wing chord, (D) maximum

wing area, (E) wing loading, and (F) aspect ratio. Circles represent medians for

each species and the black arrow points to the specimen under analysis. Source:

(20). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3.3 (A) Flight speed, (B) Horizontal accelerations, (C) Vertical accelerations. Circles

represent medians for each species and the black arrow points to the specimen

under analysis. Source: (20). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

3.4 Wingbeat period (A) scaled lower than expected under isometry. Downstroke

duration (B), Downstroke ratio (C), stroke amplitude (D), stroke plane angle (E)

and Strouhal number (F) did not change significantly with body mass. Angle

of attack increased with body size (G) as a result of a change in α1 (H), but

not from a change in α2 (I). Wing camber (J) did not change with body size,

but coefficient of lift (K) did. Circles represent medians for each species and the

black arrow points to the specimen under analysis. Source: (20). . . . . . . . . . 43

3.5 Bioinspired MAV classification depending on wingspan and mass. Source: the

author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3.6 Wing physiology. 17 markers are placed on: anterior and posterior sternum (a

and b, respectively), shoulder (c), elbow (d), wrist (e), the metacarpophalangeal

and interphalangeal joints and tips of digits III (f, g, h), IV (i, j, k), and V (l,

m, n), the hip (o), knee (p), and foot (q). Source: (21). . . . . . . . . . . . . . 44

3.7 Detailed parameters that describe wing segment morphology. a) wing segment

subdivision, b) detailed configuration of the wrist joint and attached digits when

the wing is fully extended. It shows the angles between digits that maintain

proper wing membrane tension during downstroke. Source: the author. . . . . 45

4.1 Topology. The robot has an overall of 14-DoF (not counting the 6-DoF of the

floating body). Each wing has 6-DoF and each leg 1-DoF. Source: the author. 51

x

Page 19: BaTboT: a biologically inspired flapping and morphing bat robot

LIST OF FIGURES

4.2 Detailed description of wing kinematics frames based on Denavit-Hartenberg

(DH) convention, qi corresponds to the rotation angle from axis xi−1 to xi mea-

sured about zi. The subscript i indicates the frame of reference (i = 1..6). The

inset is a top view of the right wing showing planar angles. . . . . . . . . . . . . 52

4.3 a)-b) Example of Cartesian paths during turning and forward flight, c-d) exam-

ple of wing modulation scheme for wing contraction during upstroke and wing

extension during downstroke. Source: the author. . . . . . . . . . . . . . . . . 55

4.4 Stills of wing kinematics during forward flight: a) beginning of downstroke, b)

middle downstroke, c) end of downstroke/beginning upstroke, d) middle up-

stroke. Source: (20), (22), (23), (24). . . . . . . . . . . . . . . . . . . . . . . 56

4.5 (Simulation) Wing kinematics of forward flight (both wings move symmetrically

at wingbeat frequency of f = 2.5Hz): a) Cartesian trajectories of the wrist joint

and the wingtip frame. b) Joint angles. c)-d) Joint velocities and accelerations.

For velocities, q3 = q4 (red plot) and q5 = q6 (purple plot). For accelerations,

q3 = q4, q5 = q6. Source: the author. . . . . . . . . . . . . . . . . . . . . . . . 57

4.6 (Simulation) Wing kinematics of turning flight at wingbeat frequency of f =

2.5Hz): a) Cartesian trajectories for the wrist joint and the wingtip frame of

each wing (top view). b)-c)-d) Left wing more contracted: joint angles, velocities

and accelerations. e)-f)-g) Right wing more extended: joint angles, velocities and

accelerations. For velocities, q3 = q4 (red plot) and q5 = q6 (purple plot). For

accelerations, q3 = q4, q5 = q6. Source: the author. . . . . . . . . . . . . . . . 59

4.7 Comparison between wing joint trajectories of left and right wings (q3 and q4)

considering a morphing-wing factor of fmc = 0.1. Source: the author. . . . . . 60

4.8 Rigid multibody serial chain that composes each wing. Spatial forces of each body

contain both linear fi and angular τi force components stacked into a six-dimensional

vector Fi. These forces are propagated from the wingtip {i} = n to the base frame

{0}. Subscripts R,L denote for right and left wing respectively. The resultant spatial

force (FT ) acting on the base frame {0} is the sum of spatial forces generated by both

wings. The inset shows the velocity of a rigid body i expressed in terms of ωi and vi,

and the force acting on a rigid body i expressed in terms of fi and τi. . . . . . . . . . 61

xi

Page 20: BaTboT: a biologically inspired flapping and morphing bat robot

LIST OF FIGURES

4.9 Free-body diagram of a bat in accelerating flight, indicating the aerodynamic

and gravitational forces that accelerate the center of mass (COM). Lift is per-

pendicular to the direction of flight whereas drag and thrust are parallel to the

direction of flight. The net force produced can be decomposed into net force

components parallel and perpendicular to the direction of flight (see inset). The

parallel component corresponds to the net thrust. Thus, measurements of the ac-

celeration of the COM would directly reflect the net forces acting on it. Source:

(24). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

4.10 (Experimental) aerodynamics identification –wind-tunnel–: a) the bat-robot is

mounted on top of a 6-DoF force sensor from which both lift FL and drag FD

forces are experimentally calculated as a function of the airflow speed and angle

of attack (AoA), b) Lift and drag coefficients (CL, CD) calculated from measured

lift and drag forces. Source: the author. . . . . . . . . . . . . . . . . . . . . . 67

4.11 SMA actuation configurations: a) (top) SMA joint with bias spring concept,

(medium) mechanic implementation, (below) SMA-spring model representation.

b) (top) SMA antagonistic joint, (medium) mechanic implementation that mimic

how bicep and tricep muscles operate, (below) antagonistic pair of SMAs model

representation. In both configurations SMA wires extend along the humerus

bone of BaTboT wings, acting as artificial muscles that pull the elbow joint q3.

SMA pulling forces (Fsma) produce a joint torque (τ3) that rotates the elbow,

pulling in the ”fingers” to slim the wing profile on the upstroke. Source: the

author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

4.12 SimMechanics open-loop simulator for dynamics and SMA actuation. Source:

the author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

4.13 Wing torques denoted as τi correspond to the effective forces that the each joint

i requires to rotate as defined by trajectory profiles qi (see upper inset). Torques

are with respect to the joint frames {i} assigned by modified DH convention.

The lower inset shows the model for the estimation of torques as a function of

the joint profiles. The inertial model in Algorithm 1 takes into account robot

parameters/constrains and aerodynamic loads. Source: the author. . . . . . . 77

xii

Page 21: BaTboT: a biologically inspired flapping and morphing bat robot

LIST OF FIGURES

4.14 (Simulation) a) required wing torques (τi) during forward flight. Both wings flap

symmetrically describing the wing profiles qi shown in Figure 4.5b (wingbeat

frequency f = 5Hz), b) Increments of wingbeat frequency cause the wing torques

to increase. The plot shows maximum peak values of wing torques Source: the

author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

4.15 (Simulation) a) required wing torques (τi) during turning flight: a) left wing,

b) right wing. The upper plots show the difference between the wing torque

profile that each wing requires to describe the trajectory profiles qi shown in

Figure 4.6b-e respectively (wingbeat frequency f = 5Hz). The lower plots show

how increments of wingbeat frequency cause the wing torques of each wing to

increase. Source: the author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

4.16 (Simulation) effects of wing modulation on the production of body torques and

angular accelerations at wingbeat frequency of f = 1.3Hz a) wing modulation

profile qi for the left wing (upper) and right wing (lower), b) (upper) body torques

(τθ, τφ, τψ) produced at center of mass, (medium) body angular accelerations

(θ, φ, ψ) produced at center of mass, and (lower) attitude response of the robot

(θ, φ, ψ). Values are with respect to the body frame {b}, Source: the author. . 80

4.17 (Simulation) effects of wing modulation on the production of body torques and

angular accelerations at wingbeat frequency of a) f = 2.5Hz, b) f = 5Hz, and c)

f = 10Hz. (upper) body torques,(medium) body angular accelerations, (lower)

attitude response. Source: the author. . . . . . . . . . . . . . . . . . . . . . . 81

4.18 Effects of the flapping motion of the wings on the position of the center of mass (CM)

and accelerations of the body: a) end of upstroke motion, b) end of downstroke motion.

Upper plots depicts the testbed for experimental measurements of six-dimensional in-

ertial forces (FT ) and lower plots depicts simulations for the computation of inertial

forces (FT ). Values for both measured and simulated FT are consigned in following

Figure 4.19. To conserve momentum, the body moves in opposition to the flapping

direction. During upstroke, the upward and backward acceleration caused by the flap-

ping motions of the wings produce an inertial force (red circled arrow) that moves the

body forward and downward with respect to the downstroke. This force produces a

forward-oriented component (FT,x), or inertial thrust (green solid arrow). Contrary,

during the downstroke negative inertial thrust is produced. Source: the author. . . . 83

xiii

Page 22: BaTboT: a biologically inspired flapping and morphing bat robot

LIST OF FIGURES

4.19 (Experimental) wing inertia contribution on forward and turning flight: a) simulation

model VS experimental results of pitching torques τθ (f=5Hz), b) simulation model

VS experimental results of rolling torques τφ (f=5Hz), c)-d) quantification of wing

inertia contribution into the generation of τθ and τφ at different wingbeat frequencies

f . Source: the author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

4.20 (Simulation) pitching torque response (τθ) for different positions of q2 (wings

rotated forward/backward the body) and for different wingbeat frequencies f .

The set M1 contains bio-inspired wing modulation shown in Figure 4.5b whereas

set M2 corresponds to step-input signals that rotate each joint of the wing due

to the range defined in Table 4.6 (cf. joint rotation range). Source: the author. 85

4.21 (Simulation) rolling torque response (τφ) for different positions of q2 and wing-

beat frequencies f . The set M1 contains bio-inspired wing modulation whereas

set M2 corresponds to step-input signals that rotate each joint of the wing.

Source: the author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

4.22 (Simulation) SMA phenomenological model response at different current profiles: a)

Joint rotation based on SMA strain. b) Temperatures on the SMA wire, c) Hysteresis-

loop for the nominal operation mode (Isma = 350mA), d) SMA strain VS stress. . . . . 87

5.1 Design process of BaTboT. Source: The author. . . . . . . . . . . . . . . . . . . . 91

5.2 a) weight distribution of main components, b) detailed view of main components.

Source: The author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

5.3 a) Each morphological aspect of BaTboT design has been carefully approached

to its biological counterpart, b) Cartoon of typical wingtip trajectory described

during a wingbeat cycle. Source: The author. . . . . . . . . . . . . . . . . . . . . 96

5.4 a) Pin-out and Circuit Diagrams for SMA migamotor actuator model NM706-

Super, b) actuator dimensions. Source: The author. . . . . . . . . . . . . . . . . 99

5.5 (left) half fixed-wing flapping testbed, (right) experimental quantification of

power requirements for flapping at maximum f = 10Hz. Source: The author. . . 101

5.6 Detailed flapping-wing mechanism. Source: The author. . . . . . . . . . . . . . . 102

5.7 (left) platform for wing membrane fabrication, (right) membrane issues: air bub-

bles are kept within the mixture. Source: The author. . . . . . . . . . . . . . . . 103

xiv

Page 23: BaTboT: a biologically inspired flapping and morphing bat robot

LIST OF FIGURES

5.8 Articulated-wing design: a-b) metal tendons are placed inside the bones for

connecting the elbow joint with the wrist. This enables each digit rotate as a

function of the elbow’s rotation, c) cartoon of a biological contracted wing and

its parts, d) ABS fabricated articulated wing. Source: The author. . . . . . . . . 104

5.9 a) A high camera captures the motion of each marker places along the specime’s

wings. Pictures were taking during contraction and extension of the wings, b)

detailed dimensions of wing morphology during middle downstroke. Angles and

proportions are extracted from the in-vivo experiments from (a), c) membrane-

free maximum rotation of elbow and wrist joints during wing contraction and

extension, d) maximum rotation of elbow and wrist joints during wing contrac-

tion and extension including the wing-membrane load. Source: The author. . . . 105

5.10 Antagonistic mechanism of SMA-based muscle actuators. Source: The author. . 106

5.11 Testing on the stretchable property of the fabricated silicone membrane: a) ex-

tended plagiopatagium skin, embedded tiny muscles control the membrane ten-

sion, b) the fabricated silicone membrane is attached to the wing skeleton using

Sil-Poxy RTV, c) the fabrication process described in section 5.3.2.2 results on a

light artificial skin with the required stretchable property. Also note that after

vacuum, air-bubbles are eliminated. Source: The author. . . . . . . . . . . . . . . 108

5.12 Technical overview of Dragon Skin silicone properties. Source: http://www.

smooth-on.com. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

5.13 Mid-downstroke wing camber and angle of attack are estimated as follows: (A)

A parasagittal (xg-zg) cross section of the wing was taken at the yg-value of the

wrist at the time of maximum wingspan. Six triangular sections of the wing

membrane crossed that plane and the intersections of triangle borders in the

plane (red circles) were used as estimates of membrane position. (B) The actual

curved shape of the membrane in the plane (solid black line) was estimated

using the first term of a sine series fitted to those seven points. The maximum

distance of the membrane line from the chord line (dashed grey line) was divided

by the length of the chord line to give wing camber. (C) Angle of attack (α)

was calculated as α1 + α2, where α1 is the angle of the wing chord line above

horizontal (blue dashed line), and α2 is the angle between horizontal and the

velocity vector of the wrist (red arrow) in the xg − zg plane. Source: (20). . . . . 109

5.14 Variation of camber during the wingbeat cycle. Source: (25). . . . . . . . . . . . 110

xv

Page 24: BaTboT: a biologically inspired flapping and morphing bat robot

LIST OF FIGURES

5.15 Implications of wing camber into lift and drag production (Vair = 5ms−1, fixed-

wing): a) excessive wing camber (0.48), b) proper wing camber (0.16). Source:

The author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

5.16 Micro-controller and Migamotor SMA muscle connection diagram. Source: The

author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

5.17 Comparison of attitude IMU readings before and after Kalman filtering. Source:

Arduino. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

5.18 Miga analog driver V5 pinout diagram. Source: The author. . . . . . . . . . . . . 115

5.19 Percentage of current consumption per component. Source: The author. . . . . . 116

6.1 Flight Control Architecture (FCA). Source: The author. . . . . . . . . . . . . . . 119

6.2 Experimental testbed for the characterization of SMA input power (uheating) to

output torque (τ3). Forces (F ) are measured using a force sensor with 0.318 gram-

force of resolution. Torque conversion is applied by considering the humerus bone

length (lr), as: τ3 = Flr[Nm]. This allows the identification of SMA actuation.

Source: The author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

6.3 (Experimental VS model) Bode magnitude and phase plots for NiTi 150μm SMA

Migamotor actuators. The insets show several experimental measurements of

magnitude and phase. Magnitude is given by: 20log(A/b), where A is the least-

squares estimation of the force amplitude measured using Eq. 6.3, and b is the

AC power of the input signal uheating. Phase is given by the term ϕ in Eq. 6.3.

The transfer function that fits the experimental data is shown in Eq. 6.4. This

plot also compares the model in Eq. 6.4 against the experimental data. Source:

The author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

6.4 (Above) right wing extended and contracted taking into account the load pro-

duced by the silicon membrane. (Below) Experimental measurements of SMA

output torque τ3 generated by input heating power uheating. Values are classified

by nominal and overloaded operation mode of the SMA actuators. Nominal be-

havior is achieved by applying uheating =∼ 1.36W whereas overloaded behavior

requires an input power of uheating =∼ 2.57W . Inset plots show the average peak

of produced elbow torque at both SMA operation modes. Source: The author. . 127

xvi

Page 25: BaTboT: a biologically inspired flapping and morphing bat robot

LIST OF FIGURES

6.5 Experimental measurements of joint rotation speeds of elbow joint (θ3) that are

obtained by applying heating power values at nominal (uheating = 1.36W ), and

overloaded (uheating = 3.06W ) SMA operation. Source: The author. . . . . . . . 128

6.6 (Experimental) Input Power to Output Torque small-signal response of the SMA

actuators, being uheating = a+ bsin(2πft), f = 2Hz. Source: The author. . . . . 129

6.7 (Experimental) Resistance-Motion (RM) linear relationship between SMA electrical

resistance change (Rsma) and the angular motion generated at the elbow joint (q3).

Small variations in ambient temperature (To) modify the RM relationship. The inset

shows BaTboT in the wind-tunnel. Ambient temperature has been measured with a MS

1000-CS-WC temperature sensor supplied by ATS (http://www.qats.com/Products).

Changes in Rsma are constantly measured during the experiment. Source: the author. . 130

6.8 General scheme for the inner morphing-wing control. Source: the author. . . . . . . . 132

6.9 Detailed inner loop of sliding-mode morphing-wing control with torque reference. Source:

the author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

6.10 Detailed inner loop of PID morphing-wing control with joint position reference. Source:

the author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

6.11 Typical elbow joint reference profile q3,ref during a wingbeat cycle (f = 1.25Hz). It

details how Algorithm 3 works. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

6.12 a) In-vivo recordings of a bat landing on the ceiling, cf. (26), b) closed-loop

control simulation of attitude maneuvering using the backsteping+DAF strategy.

Source: the author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138

6.13 Detailed outer and inner loops: complete description of the Flight Control Architecture

(FCA). Source: the author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

6.14 Experimental setup using the wind-tunnel of Brown University: Flight Control Archi-

tecture (FCA). Source: the author. . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

6.15 (Simulation) morphing-wing based on sliding-mode response: a) control tracking given

a sinusoidal joint trajectory (q3,ref ) at wingbeat frequency of f = 4Hz, b) phase plane

of the sliding surface upon sinusoidal input from plot (a), c) electrical current (Isma)

to drive each SMA actuator in the antagonistic configuration. Source: the author. . . . 146

xvii

Page 26: BaTboT: a biologically inspired flapping and morphing bat robot

LIST OF FIGURES

6.16 (Experimental) morphing-wing based on sliding-mode response: a) SMA output torque

τ3 given a force reference τ3,ref , b) control tracking given a bio-inspired joint trajec-

tory q3,ref at wingbeat frequency of f = 2.5Hz; it shows comparison of sliding-mode

response with and without the saturation function sat(S) within the control law, c)

electrical current (Isma) to drive each SMA actuator in the antagonistic configuration

Source: the author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147

6.17 (Simulation) morphing-wing based on PID response: a) PID control tracking given

a sinusoidal profile q3,ref (above), square profile (medium), sawtooth profile (below),

at wingbeat frequency of f = 2.5Hz, b) electrical current (Isma) to drive each SMA

actuator in the antagonistic configuration. Source: the author. . . . . . . . . . . . . . 148

6.18 (Experimental) morphing-wing based on PID response: a) measured SMA output

torque τ3, b) PID control tracking given a bio-inspired joint trajectory q3,ref , c) elec-

trical current (Isma) to drive each SMA actuator in the antagonistic configuration.

Source: the author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

6.19 (Simulation) backstepping+DAF attitude stabilization. Roll (φref ) and pitch (θref )

references are set to zero while initial attitude position is set to 20o and −20o for roll

and pitch respectively. Control efforts are also shown. Source: the author. . . . . . . . 151

6.20 (Experimental) backstepping+DAF attitude stabilization. Roll (φref ) and pitch (θref )

references are set to zero. The wind-tunnel airspeed has been set to 2ms−1 and the

controller must keep both angles close to zero. Source: the author. . . . . . . . . . . . 151

6.21 (Simulation) backstepping+DAF attitude tracking. Roll (φref ) and pitch (θref ) sinu-

soidal references are tracked. Source: the author. . . . . . . . . . . . . . . . . . . . . 152

6.22 (Experimental) backstepping+DAF attitude tracking. Roll (φ) and pitch (θ) profiles

are tracked for wind-tunnel airspeeds of 0 and 5ms−1. Tracking errors are measured.

Source: the author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

7.1 a) setup for morphing-wing testing. b) wind-tunnel setup for dynamics, aerody-

namics and control testing. Source: the author. . . . . . . . . . . . . . . . . . . . 155

xviii

Page 27: BaTboT: a biologically inspired flapping and morphing bat robot

LIST OF FIGURES

7.2 Stills of morphing-wings’ control within the wind-tunnel. The wingbeat cycle is com-

posed by two phases: downstroke and upstroke. a) Beginning of the downstroke. The

body of the specimen is lined up in a straight line, elbow joint is ∼ 58o, b) end of

downstroke, the membrane is cambered and the wings are still extended, elbow joint

is ∼ 5o, c) middle of downstroke, the wings are extended to increase lift, elbow joint is

∼ 20o, d) upstroke, the wings are folded to reduce drag, elbow joint is ∼ 45o. Figures

a-b illustrate the process to measure aerodynamics loads using the force sensor located

at the center of mass of the robot (below the body). Figures c-d illustrate the process to

measure inertial forces at the center of mass produced by both wings (no aerodynamics

loads caused by the membrane). e-f) show the beginning of downstroke and the end of

the upstroke without the membrane. . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

7.3 (Experimental) morphing-wings’ control response. a) Tracking of the elbow’s joint

trajectory at f = 2.5Hz, Vair = 0m/s i.e., no-wind. b) Close-up to a wingbeat cycle.

The two plots describe the control tracking regarding: i) Vair = 0m/s (same than plot-

a), and ii) Vair = 5m/s. c) Electrical current Isma delivered to the antagonistic SMA

actuators, and regulated by the anti-slack and anti-overload mechanisms. d) Position

tracking errors from plot-b. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158

7.4 (Experimental) Performance of the SMA actuator for longer periods of actuation: a)

Nominal operation at 1.3[Hz], b) Overloaded operation at 2.5[Hz], c) Output torque

peaks extracted from overloaded response in plot-b). . . . . . . . . . . . . . . . . . . 159

7.5 Forward flight control. Backstepping+DAF attitude tracking at: a)-b) roll and pitch

tracking with Vair = 5ms−1, c)-d) roll and pitch tracking with Vair = 2ms−1. . . . . . 161

7.6 Turning flight control. Backstepping+DAF attitude tracking at: a)-b) roll and pitch

tracking with Vair = 5ms−1, c)-d) roll and pitch tracking with Vair = 2ms−1. . . . . . 162

7.7 (Experimental) Aerodynamics measurements. a) Comparison between lift and drag

coefficients (CL, CD) with and without the motion of the morphing-wings (Vair = 5m/s,

wingbeat frequency of f = 2.5Hz). b) Lift-to-drag ratio (L/D) as a function of the

angle of attack (Vair = 5m/s, f = 2.5Hz ). c) Wind-tunnel airspeed measurements

(Vair). d) Lift (L) and drag (D) forces corresponding to plot-a (with morphing). . . . . 163

7.8 (Forward flight) benefits of the DAF to the proper modulation of wing-morphology

aimed at incrementing net forces (f = 2Hz, φref = 0o, θref = 10o): a) without the

DAF, b) with the DAF. Top: attitude tracking error and disturbance rejection; middle:

detailed wing modulation (elbow joint q3); bottom: net forces generated. . . . . . . . 165

xix

Page 28: BaTboT: a biologically inspired flapping and morphing bat robot

LIST OF FIGURES

7.9 Effects of different wing modulation profiles (no-DAF terms presented) on lift and drag

production (f = 2.5Hz, Vair = 5ms−1): a) Cartesian trajectory of the wingtip gen-

erated during a wingbeat cycle measured with respect to the base frame {0}, b) wing

modulation profile of elbow joint q3 for different backstepping parameter values of λ2

and λ4 (no-DAF-1: λ2 = λ4 = 0.1, no-DAF-2: λ2 = λ4 = 0.05), c) wing modulation

profile of wrist joint q4 corresponding to the same backstepping parameter values con-

figuration from plot-b, d) (left) net forces and (right) lift and drag coefficients generated

with the backstepping parameter configuration no-DAF-1, e) left) net forces and (right)

lift and drag coefficients generated with the backstepping parameter configuration no-

DAF-2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166

7.10 (Experimental) comparison of horizontal inertial acceleration (Axb) produced by the

wing modulation at f = 2.5Hz and Vair = 5ms−1: (above) biological data of C.

brachyotis specimen; several measurements reported in (24), (below) BaTboT; several

estimations of Axb based on measurements of inertial thrust fxb. . . . . . . . . . . . . 169

10.1 Closed-loop environment with sliding-mode morphing control. Source: the

author. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190

10.2 Closed-loop environment with PID morphing control. Source: the author. . . 191

xx

Page 29: BaTboT: a biologically inspired flapping and morphing bat robot

List of Tables

2.1 Key Glossary of bat physiology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

3.1 Description of the 27 individuals used in the study. Source: (20). . . . . . . . . . . . 40

3.2 Scaling factors for wing morphological parameters as a function of body+wing mass

mt (cf. Figure 3.2). Source: (20). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

3.3 Scaling factors for wing aerodynamics parameters as a function of body’s mass mt (cf.

Figure 3.4). Source: (20). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

3.4 Detailed wing segment geometry at downstroke (segments according to Figure 3.7a.) . . 46

3.5 Key bio-inspired geometrical parameters for modeling and design. . . . . . . . . . . . 47

3.6 Biological-based framework for modeling and design. mt = 0.125Kg . . . . . . . . . . 47

4.1 Modified Denavit-Hartenberg parameters per wing. . . . . . . . . . . . . . . . . . . . 53

4.2 (Forward flight) coefficients of the third-order polynomial curves that describe Cartesian

paths of wrist and wingtip frames fitted to the xb, yb coordinates of the body frame. . . 58

4.3 Spatial operators. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

4.4 Parameters for SMA phenomenological model . . . . . . . . . . . . . . . . . . . . . . 73

4.5 (forward flight) wing torques as a function of the wingbeat frequency f . . . . . . . . . 77

4.6 Characterization of actuators: wing torque requirements for flapping at 10Hz and

morphing1 at 2.5Hz . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

4.7 Wing and body mass influence in the generation of τθ and τφ (mt = 0.125kg) . . . . . 85

4.8 Advantage of using bio-inspired wing joint trajectories on the production of pitching

and rolling torques). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

4.9 Limits for overloading SMA operation . . . . . . . . . . . . . . . . . . . . . . . . . . 89

5.1 Main structural components of BaTboT. CAD design . . . . . . . . . . . . . . . . . . 94

5.2 Comparison of morphological parameters between the specimen and BaTboT . . . . . 95

xxi

Page 30: BaTboT: a biologically inspired flapping and morphing bat robot

LIST OF TABLES

5.3 Actuators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

5.4 Circuit pinout data for SMA migamotor actuator. . . . . . . . . . . . . . . . . . . . . 100

5.5 General values of current consumption . . . . . . . . . . . . . . . . . . . . . . . . . . 115

5.6 Fabrication costs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

6.1 Summary of SMA actuation performance. . . . . . . . . . . . . . . . . . . . . . . . . 130

6.2 List of robot’s parameters: morphological, modeling, control. . . . . . . . . . . . . . . 144

7.1 Performance data of SMA actuation for longer periods of time. . . . . . . . . . . . . . 160

7.2 Lift and drag measurements for an angle of attack of 10o and Vair = 5m/s . . . . . . . 164

7.3 List of parameters used for experiments in figure 7.8. . . . . . . . . . . . . . . . . . . 165

7.4 Summary of performance of backstepping+DAF control and its influence into wing

modulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167

xxii

Page 31: BaTboT: a biologically inspired flapping and morphing bat robot

GLOSSARY

Nomenclature

a = distance from axis zi−1 to zi measured along xi−1, (m)

Ab = extended wing area, (m2)

AoA = angle of attack, (deg)

B = extended wing length, (m)

CL, CD = lift and drag coefficients

d = distance from axis xi−1 to xi measured along zi, (m)

f = wingbeat frequency, (Hz)

i = subscript that indicates the joint frame of the wings

fi = force of body i with respect to joint frame {i}, (N)

Fi = spatial force of body i with respect to joint frame {i}FL, FD = lift and drag forces measured about body frame {b}, (N)

Fnet = net forces measured about body frame {b}, (N)

FT = propagated spatial forces of the wings with respect to the base frame {0}H = projection onto the axis of motion

Ixx, Iyy , Izz = moments of inertia of body i with respect to frame {cm} (Kgm2)

Ii,cm = spatial inertia of body i with respect to frame {cm}Ib = spatial body inertia with respect to the body frame {b}Isma = electrical current input to SMA actuators, (A)

Ji,cm = inertial tensor of body i with respect to frame {cm}lh, lr = humerus and radius bones length, (m)

Li,cm = spatial inertial moment of body i with respect to frame {cm}Mb = body and wing mass, (Kg)

ni = torque of body i with respect to joint frame {i}, (Nm)

pi,i+1 = 3x1 position vector that joins frame {i} to {i+ 1}pi,i+1 = skew symmetric matrix corresponding to the vector cross product of pi,i+1

Pi,i+1 = spatial translation from joint frame {i} to {i+ 1}Psma = input heating power to SMA actuators, (W )

qi, qi, qi = joint positions, velocities and accelerations of body i with respect to {i} , (rad, rad/s, rad/s2)

ri+1,i = 3x3 basic rotation matrix that projects frame {i+ 1} onto frame {i}Ri+1,i = spatial rotation from joint frame {i+ 1} to {i}Rsma = electrical resistance of SMA actuators, (Ω)

si,cm = 3x1 position vector that joins frame {i} to {cm}si,cm = skew symmetric matrix corresponding to the vector cross product of si,cm

Si,cm = spatial translation from joint frame {i} to {cm}Ti+1,i = 4x4 homogeneous transformation matrix that relates frame {i+ 1} with {i}u = control inputs

U = 3x3 identity operator

υi, υi = linear velocity and acceleration of body i with respect to joint frame {i}, (m/s,m/s2)

Vi, Vi = spatial velocity and acceleration of body i with respect to joint frame {i}Vb = six-dimensional body accelerations with respect to the body frame {b}α = is the angle from axis zi−1 to zi measured about xi−1, (rad)

λ = backstepping+DAF control gains

φ, θ = roll and pitch angles measured about body-frame {b}, (rad)φd, θd = desired roll and pitch angular acceleration functions (DAF), (rad/s2)

τφ, τθ = rolling and pitching torques measured about body-frame {b}, (Nm)

ωi, ωi = angular velocity and acceleration of body i with respect to joint frame {i}, (rad/s, rad/s2)

xxiii

Page 32: BaTboT: a biologically inspired flapping and morphing bat robot

GLOSSARY

Acronyms

CAD Computer-Aided Design

CM Center of Mass

DAF Desired Angular acceleration Function

DC Direct Current

DH Denavit-Hartenberg

EoM Equations of Motion

FCA Flight Control Architecture

LiPo Lithium-Polymer

MAV Micro Aerial Vehicle

MCP MetaCarpoPhalangeal

PID Proportional-Integral-Derivative

PWM Pulse Width Modulation

SMA Shape Memory Alloy

SME Shape Memory Effect

UAV Unmanned Aerial Vehicle

VTOL Vertical Take-Off and Landing

xxiv

Page 33: BaTboT: a biologically inspired flapping and morphing bat robot

1

Introduction

”Flying. Whatever any other organism has been able to do man should surely be able to do

also, though he may go a different way about it.”

– Samuel Butler.

Bats exhibit extraordinary flight capabilities that arise by virtue of a variety of unique me-

chanical features. Bats have evolved with powerful muscles that provide themorphing capability

of the wing, i.e. folding and extension of the wing during flight. To change wing morphology,

bat wings are made of flexible bones that possess independently controllable joints (24), and a

highly anisotropic wing membrane containing tiny muscles that control the membrane tension

(27). This high degree of control over the changing shape of the wing has a great impact one

the maneuverability of the animal (28), (21), (29).

In recent years the concept of morphing Micro Air Vehicles (MAVs) has gained interest

(30), (31), (32), (17). The possibility of having actuated wings has allowed the design of

new mechanisms that improve over classical fixed/rotary-wings MAV flight performance. As

a result, different morphing-wing concepts and materials have emerged together with control

methodologies that allow for accurate wing-actuation (33), (34), (35), (36), (37), (38).

The concept of morphing-wings comes from nature (39), (40). Recently, the biological

community has demonstrated a special interest in understanding and quantifying bat flight

motivated by the sophistication of their flight apparatus (24), (21), (20). The wings are highly

articulated with independently controllable joints actuated by powerful muscles that provide

the animal with a high degree of control over the changing shape of the wing during flight.

In addition, tiny muscles embedded into the highly anisotropic wing membrane contribute in

controlling the membrane tension and camber (41). Bat wings also incorporate tiny hairs that

1

Page 34: BaTboT: a biologically inspired flapping and morphing bat robot

yo

xo

zo

{o}

{b}

yb

zb

xb

roll

pitch

yaw

humerus bone

radius bone

legelbow joint

wrist jointshoulder joint

wing membrane

MCP-III

IVV

τφ τθ

Figure 1.1: BaTboT. The overall mass of the skeleton+electronics+battery is 125g. The

wingspan: 53cm (wings fully extended). Each wing of the robot has six degrees of freedom

(dof): 2-dof at shoulder, 1-dof at elbow, and 3-dof at wrist joint. The body frame {b} is a 6-dof

floating body. Rotations about the body-frame {b}-xb, yb, zb axes are designated roll, pitch and

yaw following aerodynamic conventions. Frame {o} is the inertial frame.

sense airflow conditions, and there is some evidence that this sensing apparatus contributes to

their flight efficiency (8). There is no other flying creature in nature with a similar morphing-

wing system (28), (42).

Attempting to mimic the mechanics basis of bat flight seems to have great potential to

improve the maneuverability of current micro aerial vehicles. To closely mimic the morphing-

wing mechanism of bats, muscle-like actuation seems to be an adequate solution. In this regard,

Shape Memory Alloys (SMAs) have opened new alternatives with the potential for building

lighter and smaller smart actuation systems (43), (44), (45), (46). To the best of the authors’

knowledge, there is no morphing-wing MAVs in the state-of-the-art with highly articulated

wings inspired by the biomechanics of bats, actually the only works attempting to reproduce

bio-inspired bat flight using SMAs are presented in (10) and (47). Most of the experiments

in (10) were carried out with only a two degree of freedom wing capable of flapping at 3Hz.

Despite the fact that their robot is able to achieve accurate bio-inspired trajectories, the results

presented lack experimental evidence of aerodynamics measurements that might demonstrate

the viability of their proposed design. Moreover, neither (10) or (47) detail how to control the

SMAs for achieving the bio-inspired motion of the wings. Other works have also explored how

2

Page 35: BaTboT: a biologically inspired flapping and morphing bat robot

1.1 The problem and motivations

to optimize aircraft performance based on the aerodynamics of bat wings (11), (48).

Motivated by the potential behind bat flight and the lack of highly articulated morphing-

wing MAVs (not necessarily bat-like), this thesis presents a novel bat-like micro air vehicle

inspired by the morphing-wing mechanism of bats: BaTboT (cf. Figure 1.1). This thesis is

about:

The design and fabrication of the first highly articulated morphing-wing bat-like

robot. A novel strategy for the flight control will allow BaTboT to efficiently

maneuver by means of modulating wing inertia, without the need of any extra

mechanism such as ailerons, rudders, or back tails.

1.1 The problem and motivations

”Bats, the mysterious nocturnal mammals that are guided by sound, might hold the secret to

more-efficient flying machines.”

The problem

Morphing-wing aircrafts have emerged as a direction to enhance the efficiency of flight by

changing the wing profile. There is growing interest in the energy cost of flight and learning

from nature is the key to optimize efficiency. However, nature flyers such as insects, birds

or bats have extreme complexity in their flight apparatus and attempting to mimic part of

that complexity using artificial counterparts presents several challenges. Among these flyers

bats have evolved with truly extraordinary aerodynamic capabilities that enable them to fly

in dense swarms, to avoid obstacles, and to fly with such agility that they can catch prey

on the wing, maneuver through thick spaces and make high speed 180o degree turns. More

importantly, biologists have discovered that by flexing their wings inward to their bodies on the

upstroke, bats use only 65 percent of the inertial energy they would expend if they kept their

wings fully outstretched. Unlike insects and birds, bats have heavy muscular wings with hand-

like bendable joints and it is precisely this higher degree of dexterity that allows bats to save

energy during flight than any other flying creature.

The main problem to tackle in this thesis is how to optimize efficiency in terms

of net force production (inertial and aerodynamical) by developing a novel micro

3

Page 36: BaTboT: a biologically inspired flapping and morphing bat robot

1.1 The problem and motivations

aerial vehicle prototype with unprecedent highly articulated morphing-wings

inspired by the bat flight apparatus.

Solving this problem can help further the continued development in small unmanned flying

vehicles that waste minimum energy at the expense of incrementing payload capacity.

The hypothesis

In nature bats modulate wing inertia to improve on dynamics and aerodynamics response.

Based on this biological fact the following question is formulated:

Could a micro aerial vehicle inspired by the biomechanics of bats take advantage of the high

dexterity provided by the morphing-wings aimed at improving flight efficiency?

To this purpose, the following hypothesis is proposed:

Quantifying the effects of wing inertia in terms of thrust and lift production and

therefore including wing inertia information into the flight controller will allow

for the proper modulation of wing kinematics that finally would produce and

increase of net forces, thereby improving on flight efficiency.

Motivations: learning from bats

Bats can carry up to 50 percent of their weight and execute airborne maneuvers that would

make a bird or plane fall out of the sky. Bats use sophisticated echolocation to navigate, but

on top of that, hundreds of tiny hair sensors on the wing membrane that feed flight data are

used by the animal to improve on aerodynamics performance. Non other flying specie resemble

the way how bats sense airflow and adjust the wings to improve on fight.

Wing mass is important and it is normally not considered in flight. In bats, there is biological

evidence that the inertial forces produced by the wings have a significant contribution into the

attitude movements of the animal, even more significant than aerodynamic forces (24), (20).

In fact, bats perform complex aerial rotations by modulating solely wing inertia

(26), (23). This means bats are able to change the orientation of the body during

flight without relying on aerodynamic forces and instead by changing the mass

distribution of its body and wings. Inertial forces are likely to be significant in bats

4

Page 37: BaTboT: a biologically inspired flapping and morphing bat robot

1.1 The problem and motivations

a) b)

c) d)

VTOL

Figure 1.2: a) bats are agile flyer for hunting preys in air, water or even ground, b) bats can hover

like hummingbirds, consuming less energy (10 − 20Hz of flapping frequency), c) bat wings can

camber, stretch, extend and fold like no other flying animal in nature, d) VTOL flight capacity,

Source: Breuer Lab, http://brown.edu/Research/Breuer-Lab/research/batflight.html

because the mass of the wings comprises a significant portion of total body mass,

ranging from 11% to 33% and because wings undergo large accelerations (49).

Unlike birds or insects, whose wings are comparatively rigid and lighter, bats have wings

with more than two dozen independent joints, much like a human hand. This allows them to

manipulate the thin, flexible membrane that covers the wings in ways that can generate more

lift or greatly reduce drag. Surfaces of bat wings also curve more than a bird’s (camber) –

providing greater lift for less energy – while their extraordinary flexibility allows them to make

a 180-degree turn in less than half their wingspan, a radius impossible for any bird or existing

plane. Alike hummingbirds bats also can hover and power VTOL flight (vertical take-off and

landing) but with the main difference of saving energy due to the wingbeat frequency. Figure

1.2 shows some stills of amazing maneuvers bats are able to perform during flight.

The review of specialized literature reveals that many evolutionary biologist are attempting

to unlock the secrets of bat flight. Nonetheless, from a robotics perspective, lack of research

could be due to the complexity of the flight apparatus and the challenges to mimic part of that

complexity using an artificial counterpart.

5

Page 38: BaTboT: a biologically inspired flapping and morphing bat robot

1.1 The problem and motivations

Bats exhibit extraordinary flight abilities due to their unique wing structure which is quite

different then in other flyers like birds and insects. The flexible, flapping wings of bats may

pave the way toward versatile new types of aircraft, evolutionary biologist at Brown University

have revealed. 1 The sophisticated analysis looked at bats in a wind tunnel to uncover the key

differences between how mammals and birds stay aloft. Bats turn out to have a high degree of

control over the changing shape of their wings. These mammals can therefore generate lift as

their wings move both up and down; a big advantage when hovering.

Also aerodynamic forces generated by bat wings during flight are far more complicated

than those of birds. Bird wings operate almost as if they were airplane wings on hinges. By

comparison bat wings are more flexible and the material of the skin and bones are more stretchy

(the bones actually bend when the bat is flying). At slower speeds the morphing wings of bats

seem to have advantages in terms of maneuverability and energy savings, key issues that would

make bat-like aircrafts superior to bird-like or conventional aircrafts, especially in search-and-

rescue operations and covert surveillance.

To change wing morphology bat wings have an extremely high degree of articulation compris-

ing the elbow, wrist and finger joints which makes it more feasible to reproduce its mechanical

parameters using existing light materials. Its tendons and muscles are much smaller than those

found in birds and thus easier to model. With a thickness that varies in the range of 0.04 to

0.15mm, the wing membrane consists of many elastic veins and tiny muscle fibers that allow

the wing to be extended, folded and cambered. The skin of the membrane is actually very stiff,

its elasticity relying upon the fine and wrinkle texture that flattens out to create a taught airfoil

when the wing is extended.

The advent of the smart materials made possible the design of light wings that mimic the

complex motion of the morphing flight. E.g, the elbow and the wrist joints can be reproduced

such that the wing can fold and rotate its different segments minimizing the drag and the

negative trust in the upstroke motion. Artificial muscles designed from smart materials actuate

these motions during both downstroke and upstroke portions of the wingbeat cycle. By using

shape memory alloys (SMAs) as muscle like actuators behaving as biceps and triceps along the

wing-skeleton structure of the robot, the wings can extend and contract under the control of

the SMA wires that switch between two shapes when different currents are applied. The wires,

between the ”shoulder” and ”elbow” of the robot, rotate the elbow, pulling in the ”fingers” to

1Part of this research has been done with the collaboration of the Swartz Lab and Fluid Mechanics Laboratory

at Brown University.

6

Page 39: BaTboT: a biologically inspired flapping and morphing bat robot

1.2 Objectives

slim the wing profile on the upstroke. SMA might enable the contraction and extension motions

the wings in a similar way to the biological counterpart but more important, it can provide the

design of light actuated morphing-wings.

1.2 Objectives

This thesis presents a novel bat-like Micro Aerial Vehicle BaTboT with actuated morphing

wings that can be efficiently modulated by a novel flight controller that uses wing inertia

information to that purpose. The goal, to improve on flight performance in terms of lift/thrust

production and drag reduction.

In brief, this thesis provides both theoretical and experimental foundations for designing

highly articulated morphing-wing MAVs aimed at enhancing flight performance via proper wing

modulation (not necessarily bat-like).

The specific objectives of this research are:

1. To analyze and select which bat-specie would be suitable to be mimicked by an artificial

counterpart.

2. To study the mechanistic basis of bat flight. Based on published biological data that

unveils key aspects of bat’s morphology, physiology and aerodynamics performance, to

define a biological-based framework useful for designing a bio-inspired bat-like robot.

3. To formulate mathematical models for: i) body and wing kinematics, ii) dynamics (iner-

tial contribution), iii) aerodynamics (lift and drag production), and iv) SMA muscle-like

actuation.

4. To validate mathematical models against experimental data.

5. To design and fabricate BaTboT using the proposed biological-based framework.

6. To formulate a morphing-wing controller for the proper regulation of SMA actuators.

7. To formulate an attitude controller for the proper morphing-wing modulation that pro-

duces forward and turning flight.

8. To analyze and discuss the performance of BaTboT in terms of: i) accurate and fast SMA

actuation of morphing-wings, ii) inertial contribution on thrust production, iii) lift-to-drag

ratio, and iv) power consumption.

9. To demonstrate BaTboT would be capable of forward and turning flight via wind tunnel

testing without the need of external appendices such as rudders, ailerons, propellers, etc.

10. To discuss about the potential of the proposed methodologies towards real flight.

7

Page 40: BaTboT: a biologically inspired flapping and morphing bat robot

1.3 Methods

1. Biological study of bat flight 2. Bio-inspired design criteria

3. Methods: modeling + simulation

4. BaTboT fabrication 5. Control (measurements)

6. Experimental results

SMA actuation

dynamics(wing modulation)

aerodynamics

Figure 1.3: Structural steps to be followed during the thesis aimed at the development of BaT-

boT. The pictures depicted herein, correspond to the final BaTboT prototype. The forthcoming

chapters will introduce each step with all the details. Source: The author.

1.3 Methods

This subsection briefly summarizes the methods used for the development of BaTboT. The

following procedures will be approached aimed at achieving the main goals of this thesis. Figure

1.3 graphically details these procedures which will be all cover within each chapters of this

document.

1. Biological study of bat flight:

It presents a detailed study of the most relevant issues that describe bat flight: i) bio-

mechanics, ii) morphology, iii) physiology, iv) muscle actuation, v) kinematics, and vi)

aerodynamics performance. This study has been based on the most specialized biological

literature review from (50), (51), (52), (22), (23), (24), (20).

8

Page 41: BaTboT: a biologically inspired flapping and morphing bat robot

1.3 Methods

2. Bio-inspired design criteria:

It quantifies key design criteria based on the studied biological data analyzed in the

previous procedure. It summarizes these criteria into three fields: i) morphology, ii)

kinematics and iii) aerodynamics. Most relevant morphological parameters are: wingspan,

body and wing mass, wing area and wing-bone lengths. Kinematics parameters are:

wingbeat frequency and wingstroke trajectories. Aerodynamics parameters are: lift and

drag forces, wing membrane camber and angle of attack.

3. Modeling:

It defines morphology and kinematics frameworks of BaTboT. Wing kinematics are for-

mulated using modified Denavit-Hartenberg (DH) convention frames(53), whereas body

kinematics is designated by roll, pitch and yaw motions with respect to the body-frame

(following aerodynamic conventions (54)). Basic rotation matrix are used to express how

kinematics variables are propagated from the wings to the body. This allows for the

formulation of an integrated inertial model that mainly consists on: i) Newton-Euler

dynamics equations of motion expressed by spatial algebra notation (55), and ii) SMA

thermo-mechanical actuation based on existing phenomenological models that describe

the shape memory effect (56). Here, SMA performance is quantified in order to assess

the limits of this actuation technology. Also, the influence of wing inertia on robot’s

maneuverability is analyzed using the inertial model.

4. Design-Fabrication:

It approaches the design/fabrication problem. It shows a detailed description for the bio-

inspired development of: i) body and wing skeleton, ii) wing membrane, iii) actuation

mechanisms, and iv) hardware components.

5. Control (measurements):

It tackles the control problem. Two control layers are developed: i) morphing-wing

controller and ii) attitude controller. The former regulates the amount of input heating

power to be delivered to the SMA muscles. SMAs actuate to change the shape of the

wings (contraction/extension). The latter drives the former. It regulates the attitude

motion of the robot (roll y pitch) by means of proper wing modulation.

The novelty of the attitude control strategy is due to the incorporation of wing iner-

tia information within the control strategy. The idea behind this approach is aimed at

improving the attitude response of the bat-like MAV. The proposed controller is called

baskstepping+DAF (DAF, desired angular acceleration function). Such enhancement is

9

Page 42: BaTboT: a biologically inspired flapping and morphing bat robot

1.4 Original Contributions of this Work

based on the assumption (motivated by the cited biological studies) that bats efficiently

generate forward thrust by means of inertia wing modulation, taking advantage of relevant

wing-to-body mass ratio.

6. Experimental results:

It concludes with experiments aimed at:

• assessing the performance of the SMA muscles driven by the morphing-wing con-

trol. Performance will be quantified in terms of actuation speed, output torque and

fatigue,

• evaluating the accuracy of the attitude controller for tracking pitch and roll references

under the presence of external disturbances caused by aerodynamics loads at high

airspeeds,

• demonstrating the assumption of incrementing net body forces thanks to the wing

inertial modulation driven by the attitude controller, and

• showing the potential of the proposed methodologies toward achieving the first bat-

like MAV capable of autonomous high maneuverable flight.

1.4 Original Contributions of this Work

The original contributions of this work cover four areas:

Bio-inspired design and modeling

I BaTboT is the first Micro Aerial Vehicle composed by highly articulated wings, with 12

degrees of freedom counting both wings. No other platforms in the literature have a similar

amount of joints in their wings-system.

II The design process of BaTboT has been entirely conceived based on a comprehensive

analysis of biological data. The data from in-vivo experiments reported in the specialized

literature allowed for the definition of a bio-inspired design framework which defines every

aspect related to morphology, kinematics, and aerodynamics as a function of body and

wing mass.

III BaTboT is the first bio-inspired Micro Aerial Vehicle capable of maneuvering just by using

biomechanics parts found in nature, such as: wings, legs, the body, the wings’ membrane,

10

Page 43: BaTboT: a biologically inspired flapping and morphing bat robot

1.5 Thesis outline

etc. Most of the concepts found in the literature make use of extra mechanical parts that

are based on avionics principles, such as: ailerons, propellers, rubbers, back tails, etc.

IV An inertial model aimed at studying the influence of wing inertia on the production of net

forces for maneuvering.

SMA actuation and power

V Identified linear models for a ©Migamotor SMA actuator relating output torque with

input power.

VI Quantification of SMA limitations in terms of fatigue and actuation speed. It defines

the trade-off between input power, output torque, and actuation speed. This trade-off is

essential for the designing process of SMA muscle-like actuation mechanisms in similar

applications.

VII Accurate and faster position control of the SMAs (up to 2.5Hz in actuation speed) thanks

to re-adapted anti-overload and anti-slack mechanisms from (9). Normal rates of SMA

actuation speed range between 1 − 2Hz. It also uses SMAs as sensors, saving on weight

and energy.

Attitude control

VIII An enhanced backstepping control law denoted as backstepping+DAF. The Desired angular

Acceleration Function (DAF) incorporates wing inertia information aimed at the proper

modulation of the wings. This improves on attitude tracking and increments the production

of inertial thrust and reduces drag forces.

1.5 Thesis outline

Each chapter of this thesis beings with a General Overview about the problems to be addressed

and a brief description of the methods to be introduced. Thereby the end of each chapter con-

cludes with brief remarks about the topics presented. This document is organized as follows:

Chapter 2 is about a Literature Review. State-of-the-art research is introduced from

specialized literature covering areas such as: i) Nature flyers with a focus on bat biology, ii)

11

Page 44: BaTboT: a biologically inspired flapping and morphing bat robot

1.5 Thesis outline

Shape Memory Alloys as an alternative for actuation, and iii) Current Micro Aerial Robotic

platforms.

Chapter 3 is about biological inspiration. Key parameters of biomechanics basis of bat

flight are studied and unified into a bio-inspired framework for robot design. Relevant biological

data is also highlighted aimed at the proper formulation of robot’s models.

Chapter 4 is about Modeling. Based on biological data, BaTboT’s morphology, kine-

matics, dynamics, identified aerodynamics, and SMA actuation are defined and modeled using

mathematical frameworks. Basic maneuvers are defined by showing the influence of wing inertia

modulation on robot’s maneuverability.

Chapter 5 is about Design. Here, the design process and fabrication of BaTboT’s compo-

nents are introduced. It shows novel approaches for bio-inspired design and for the development

of low-mass high-power circuits.

Chapter 6 is about Control. It presents novel control techniques to: i) approach faster SMA

morphing-wing modulation and ii) enhance attitude regulation that allows for more efficient

flight control.

Chapter 7 presents the experimental tests carried out. Experiments are conducted to

demonstrate: i) morphing-wings control accuracy and speed, ii) SMA performance, iii) aerody-

namics performance, and iv) overall flight control.

Chapter 8 concludes the thesis with important remarks on the obtained results. Conclu-

sions are focused on the ares of: i) bio-inspired bat design, ii) SMA as muscle-like actuators,

iii) BaTboT’s overall control, and iv) General performance of the platform.

12

Page 45: BaTboT: a biologically inspired flapping and morphing bat robot

2

Literature Review

I may be biased, but I think bat flight is inherently fascinating.

2.1 General Overview

The knowledge from mechanisms of the biological species is the key input for designing intelli-

gent systems and exploring new technologies that allow to mimic key functionalities that nature

has perfected during millions of years of evolution (50), (24).

This chapter presents a comprehensive review on key information related to the physiologi-

cal and morphological functions of biological bats, aimed at providing an integrated view of the

structural design of a bat-like robot. A specialized literature survey from (57), (58), (51), (59),

(60), (61), and (1) will provide significant insights into the requirements to mimic key function-

alities of biological flight, with an special emphasis on attempting to replicate the mechanistic

basis of the bat flight apparatus using smart materials (62), (43), concretely, Shape Memory

Alloys (SMA) acting as artificial muscles that can be controlled to provide high dexterity of

wing modulation (morphing-wings).

This chapter reviews the state-of-the-art in relation to:

• Nature flyers: comparison of some aspects of bat biomechanics with those from birds and

insects.

• Bat flight research: biological insights on quantifying bat flight.

• Shape Memory Alloys (SMA) for muscle-like actuation.

• Morphing-wing MAVs that use smart actuators.

13

Page 46: BaTboT: a biologically inspired flapping and morphing bat robot

2.2 Nature flyers

2.2 Nature flyers

The variety of flying animals is staggering, ranging from mosquitoes to hummingbirds, birds,

flying foxes, eagles. And yet they all rely on the same physical processes and almost the same

mechanism of locomotion. Nonetheless, key variations of that mechanism through evolution

has defined different patterns of flight, specialized according to the environment of each flying

specie. This section highlights some of the difference between nature flyers, emphasizing into

the unique mechanism of the bat flight apparatus.

2.2.1 Biomechanics: insects, birds, and bats

Four groups of animals have evolved powered flight: i) insects, ii) pterosaurs, iii) birds, and

iv) bats. Though these groups evolved powered flight independently, all use roughly the same

flapping pattern (cf. Figure 2.1). The wings of the three vertebrate groups (pterosaurs, birds

and bats) all evolved from a modified forelimb. Pterosaurs, extinct relatives of dinosaurs, had

wings supported mostly for a single, enormously elongated fourth finger and a wing membrane

stretched from the arm/hand skeleton to the side of the body.

Butterfly Draganfly

Cockroach Beetle

FlyTrue bug

Bats

Birds

pterosaurs

Figure 2.1: Structure of animal wings showing the main skeletal support. (Left) vertebrates,

(Right) insects.

Among living vertebrates, bats are the only mammals able to fly. Their wings are su-

perficially similar to Pterosaurs wings in that they have greatly elongated fingers supporting

a flexible membrane, but their anatomy is actually quite different mainly because the arm

skeleton (upper arm and forearm) supports much larger proportion of the wing in bats than

14

Page 47: BaTboT: a biologically inspired flapping and morphing bat robot

2.2 Nature flyers

Pterosaurs. In addition bats can adjust its four wing fingers independently, gaining more level

of flight control.

On the other hand, birds have a different wing-system, mainly because most of the surface

of their wings consists of feathers. The arm/hand skeleton extends barely half the length of the

wind, and the fingers are reduced to little more than three short spurs of bone.

Insects were the first animal group to evolve flight, and they are the only invertebrates to

have done so. Insects wings are fundamentally different compared to vertebrates in several

ways. Vertebrate wings are all modified legs (forelimbs), while insect wings evolved separately

from their legs. Insect legs are attached to the bottom of the thorax and the wings are attached

to the upper side of the thorax. Flying vertebrates have muscles as well as bones with joints

out in the wing, so vertebrates flyers can directly control the shape and movements of their

wings. This functionality of changing wing shape is also known as morphing wings. In insects,

the membrane contains embedded tiny veins that carry blood vessels. These veins are anchored

to a complex set of tiny skeletal structures, -axillary scleritis-, that make up the wing hinge. By

adjusting the position of these structures, the insects can push or pull on different veins, which

can, in turn, adjust the shape of the wing. However, due to their anatomy, that kind of wing

morphology is limited and they cannot flex their wings or bend them to shorten their span.

In that regard, bat morphing wings are more sophisticated, with an incredible potential for

high maneuverability flying at lower speeds compared to birds. To facilitate the choice of an

optimal type of natural wing as a model for the BaTboT design, a number of small natural

fliers are thoroughly analyzed from specialized literature, and their morphological and flight

parameters are compared particularly in terms of the wingspan, wingbeat frequency, the wing

loading, lift-drag, and power.

Wingspan (S) and Aspect Ratio (AR) The wingspan (cf. Figure 2.2a) is perhaps the

most important morphological measurement required on a flyer, after body mass. It is the

distance between the two wing tips when the wing is fully extended during the downstroke:

S = 2lm + 2B, (2.1)

where 2lm is the body width of the flyer and B is the mean wing length. A longer wingspan

is generally more efficient because the flyer suffers less induced drag and its wingtip vortices

do not affect the wing as much. However, the long wings mean that the flyer has a greater

moment of inertia about its longitudinal axis and therefore cannot roll as quickly and is less

15

Page 48: BaTboT: a biologically inspired flapping and morphing bat robot

2.2 Nature flyers

a) b)

d)c)

Figure 2.2: Comparison for several species of small-scale flyers: a) Wingspan, b) Aspect ratio,

c) Wing loading, d) Wingbeat frequency. Source from (1), (2), (3).

maneuverable. The aspect ratio AR (cf. Figure 2.2b) is a powerful indicator of the general

performance of a wing. It describes the shape of the wing by means of the ratio between the

wingspan and the wing’s mean chord, as defined in Eq. (2.2). High aspect ratio wings are

associated with a reduced power cost to create comparable lift. The species that use a gliding

flight, like condors, eagles exhibit a high aspect ratio that allows them to stay aloft for large

periods of time using the least amount of energy.

AR = S2A−1b , (2.2)

Where S is the wingspan and Ab the mean wing area. The damselflies tend to have the

highest aspect ratio between the studied species, making them suitable for gliding flight with a

reduced amount of energy lost through wingtip vortices. In general, bats present relatively low

aspect ratios indicating a higher maneuverability with a loss of performance due to a higher

induced drag.

Wing Loading

Wing loading (WL) refers to the ratio of the body mass to the wing surface area of the

flyer. This measurement relates the flight speed with the flyer’s maneuverability. Lower wing

16

Page 49: BaTboT: a biologically inspired flapping and morphing bat robot

2.2 Nature flyers

loading allows slower flight (still producing necessary lift), whereas larger wing loading makes

the animal to fly faster.

WL =MbA−1b , (2.3)

Maneuverability is also dependent on wing loading since the minimum radius of turn is

proportional to body mass. Comparing the wing loading in Figure 2.2c, (24) has shown that

bats are able to achieve turns of 180deg in just three wingbeats using an intricate combination

of banked and crabbed mechanisms. By studying the wing kinematics during these turns, (23)

has shown that bats can maneuver in narrow spaces less than half of their wingspan.

Wingbeat Frequency (f) is one of the most interesting flight parameter for flapping flight

and is affected by the body mass, wingspan, wing area and the wing moment of inertia. Using

a combination of multiple regressions and a dimensional analysis, (63) experimentally derived

that the larger the wingspan S, the smaller is the wingbeat frequency f to keep the flyer aloft,

and that an increase of the body mass, should be balanced by an increase wingspan S and

in wing area Ab in order to reduce the wingbeat frequency. Figure 2.2c shows the average

comparison of wingbeat frequencies for several flyer species.

Lift, Drag, and Reynolds number

Lift is defined as the component of force orthogonal to flow and so perpendicular to drag

(cf. Figure 2.3a). The lift and drag forces of the wing are defined relative to the direction of

the airflow over the wing, not relative to the directions in which the animal flies or gravity acts.

As a rule, lift is perpendicular and drag is parallel to the direction of the airflow. Lift and drag

have a resultant force, which is tilted forward during the downstroke and because of this it has

a vertical component, the upward force, and a horizontal component that is tilted forward, and

so it is providing the thrust.

The total lift force FL can be expressed through definition of a lift coefficient CL. Likewise,

the total drag FD acting on a flier is the sum of the effects of pressure and viscous drag. The

comparison of drag of different fliers is facilitated through the definition of a non-dimensional

drag coefficient CD. Both lift and drag are calculated as:

FL = 0.5CLρAbV2air

FD = 0.5CDρAbV2air

(2.4)

17

Page 50: BaTboT: a biologically inspired flapping and morphing bat robot

2.2 Nature flyers

a) b)

c) d)

Figure 2.3: a) Cartoon of aerodynamics forces acting on a typical wing design, b) Lift and drag

changes with the angle of attack for a typical wing design. c) Comparison of the lift to drag ratio.

d) Power Flight. Source from (1), (2), (3).

The term Ab is the mean wing area of the flyer and Vair os the airspeed. Both lift and drag

forces acting together with inertial forces are related by the Reynolds number Re. In other

words, the formulation of the Reynolds number is given by the ratio of inertial of inertial to

viscous forces varies with the ratio of the fluid density and the fluid viscosity (2). Most bat

specimens range from a Re between 103 and 104. Also, another important characteristic of a

given airfoil is the lift to drag ratio, L/D (cf. Figure 2.3c). The higher the lift to drag ratio

of the airfoil, the less thrust is needed to produce that required lift. Furthermore, bats use 35

percent less energy by reducing aerodynamic drag (cf. power-to-mass ratio in Figure 2.3d).

2.2.2 Bat biology

Bats are mammals, which means that they are warm-blooded, have fur, and produce milk. Bats

belong to the Order Chiroptera within the class Mammalia of the kingdom Animalia. Within

the Order Chiroptera bats can be divided into two groups depending on its taxa: Microchi-

roptera and Megachiroptera. Microchiroptera (or microbats) in general navigate by echolocation

generating ultrasound via the larynx that emits the sound through the nose or the open mouth.

Microbats call range in frequency from 14k to over 100k hertz, well beyond the range of the

human ear (typical human hearing range is considered to be from 20Hz to 20kHz). Microbats

18

Page 51: BaTboT: a biologically inspired flapping and morphing bat robot

2.2 Nature flyers

are 4 to 16cm long. They usually have short faces, well-developed tails, and do not have a claw

on their second finger. Megachiroptera (or megabats) are ”fruit-bats” that generally navigate

by sight (although a few use echolocation). In contrast to Microchiroptera, these bats have a

longer face, and a claw on their second finger. The largest (Pteropus vampyrus) reach 40cm

in length and attain a wingspan of 182cm, weighing in at nearly 1kg-f. The smallest bat is

the Craseonycteris thonglongyai, an insect-eating bat that has a wingspan of only 0.15m and

weighs few gram-force.

Glossary and terminology

Table 2.1: Key Glossary of bat physiology

Physiology Terminology

Membrane Skin

Patagium: is the wing membrane having an average thickness of about 0.03mm. It is

composed by elastic fibers and bundles of muscle fibers.

Plagiopatagium: is the large portion of the wing, between body and fifth digit. This is

the lift-generating section and its camber is controlled by flexing the body axis and/or the

fifth digit.

Propatagium: is the small portion between the shoulder, elbow and Carpus. It is also

used specially to produce lift.

Dactylopatagiums: is the digit membrane and is especially a propelling or thrust gen-

erating portion. Its camber can be changed by flexing the digits.

Main Bones

Scapula: provides attachment for many of the flight muscles and is mobile across the back

of the rib cage during the wing beat cycle

Clavicle: enables a wider arc of rotation for the humerus. During the upstroke the scapula

slides back to its more dorsal position.

Humerus: bone that connects the shoulder and the elbow joints.

Radius: bone that connects the elbow and the Carpus joints.

Third MetaCarpoPhalangeal (MCP-III): digit bone that connects the wrist/carpus

joint with the wingtip point.

Main joints

Shoulder: Complex 3-DoF joint formed by three bones: scapula, clavicle and humerus.

It is useful for producing flapping motion.

Elbow: 2-DoF joint useful for achieving wing morphing capacity. 1-DoF produces the

forearm (radius) flexion and the second one that produces the roll of the wrist resulting in

a leading edge flap.

Wrist/Carpus: Complex 3-DoF joint capable of achieving roll, pitch and yaw motion

(similarly to humans wrist). Roll helps in generating more thrust, pitch helps in flexing

the palm to approach the forearm in a downward and upward direction, and yaw helps in

flexing the forearm in a backward direction.

Main muscles

Pectoralis major, subscapularis, serratus anterior, clavodeltoideus: These are the

main four muscles involved during the wing downstroke cycle.

Trapezius, costo-spino-scapular, deltoid: These are the main three muscles involved

during the wing upstroke cycle.

Supraspinatus, triceps, biceps, occipito-pollicalis, coraco-cutaneus, humeropata-

gialis, tensor plagiopatagii: These are the group of muscles that control the flight

membranes.

Since a large part of terminology has a biological provenience, it is useful (from an engineer-

ing point of view) to introduce and explain some of the most frequent terms used for describing

bat physiology (64). These terms are consigned in Table 2.1. The forthcoming subsections will

19

Page 52: BaTboT: a biologically inspired flapping and morphing bat robot

2.2 Nature flyers

Figure 2.4: Bat anatomy. A) body anatomy, B) Pectoral skeleton, C) Dorsal view of retracted

right arm, D) Dorsal view of expanded right arm. Source (4).

graphically show and explore bat physiology in more detail.

Bat anatomy

In terms of anatomy bat wings are built on the basic pattern of a mammalian limb. It is

an analogous structure to the human arm and hand (in fact, the name ”chiroptera” is Greek

for ”handwing”), but the relative sizes of most bones and muscles are very different. The bat

also has unique muscles in the patagium, chest and back, to power the wing during flight.

The wing consists of the upper arm, forearm, wrist and hand. The bones of the hand and

the four fingers are greatly elongated, light and slender to provide support and manipulate

the wing membrane, called the patagium (cf. Figure 2.4). The wings of bats are much thin-

ner than those of birds, so bats can maneuver more quickly and more accurately than birds (63).

Bat flight muscles

Bats use the pectoral muscle as the main adductor or depressor, which means that draws the

wing toward the median line of the body (52). In (65) Prof. Norberg studied the direction of pull

of the main flight muscles in bats, given an insight about how wings move during downstroke

and upstroke cycles of the wingbeat.

20

Page 53: BaTboT: a biologically inspired flapping and morphing bat robot

2.2 Nature flyers

Pectoralis Subscapularis

Serratusventralis

Depressors (Adductors)

Levators (Abductors)

Rhomboideus

Spinotrapezius

Spinodeltoideus

ClavodeltoideusSupraspinatusAcromiotrapeziusAcromiodeltoideus

Downstroke UpstrokeM. subscapularis

ScapulaM. trapezius

M. rhomboideusM. spinodeltoideus

M. acromiodeltoideus

Clavicle

SternumM. pectoralis

M. serratus

M. clavo-deltoideus

BAT stroke muscles

Figure 2.5: Bat muscle structure to power wingstroke motion.

As shown in cf. Figure 2.5, bats have four downstroke muscles: pectoralis major, subscapu-

laris, part of serratus anterior, and part of the deltoid . Together, these muscles constitute

about 12% of the bat’s body weight. The upstroke is powered when required by the remainder

of deltoid, trapezius, the rhomboids, infraspinatus, and supraspinatus (costo-spino-scapular).

The scapula provides attachment for many of the flight muscles and is mobile across the back

of the rib cage during the wingbeat cycle.

The first digit of the wings is small and clawed and it is used during climbing and walking.

The muscle extensor carpi-radialis muscle inserts on the base of the metacarpal of the second

digit. This muscle pulls forward on the metacarpal and a ligament between digits II and III

transfers tension to the third digit, thus keeping the outer part of the wing taut and extended.

Digits IV and V extend across the chord of the wing. The muscle abductor along the ventral

surface of the 5th digit can bring about changes in the camber of the membrane (cf. Figure

2.5).

In this thesis, bats’ physiology, specially the muscles, are fundamental for the understand-

ing of maneuverability and efficiency during flight. In addition, wing kinematics will be also

derived from the analysis of how muscles move the wing bone structure and membrane in order

to generate the wing stroke patterns.

21

Page 54: BaTboT: a biologically inspired flapping and morphing bat robot

2.2 Nature flyers

M coraco - cutaneusPropatagium

M occipito - pollicalisM humeropatagialis

Chiropatagium

Plagiopatagium

M bicepsM triceps

M tensor plagiopatagiiM extensor carpi radialis

M extensor carpi ulnaris

Embedded tiny muscleswithin the membrane

Figure 2.6: Bat wing membrane

Wing Membrane

The proximal segment of the wing (plagiopatagium and propatagium) produces most of the

lift, and the distal segment, chiropatagium, produces most of the thrust during the wingbeat

cycle (cf. Figure 2.6). The proximal segment is maintained at an appropriate camber and angle

of attack for the production of lift throughout the wingbeat cycle. The degree of curvature of

the fifth digit has an important role in determining and maintaining the angle of attack and

the camber of this segment. By retaining its curvature and angle of attack throughout the wing

beat cycle, the fifth digit partly controls the lift produced by the proximal segment (58).

When the bat is at rest its wings are folded in an accordion shape-like. Then, at the

beginning of the flight, a rapid extension of the fingers of the wings occurs, extension that is

generated by the contraction of a single muscle: the supraspinatus.

The opening speed of bat wings is very rapid due to its unique folding and unfolding patterns

(morphing-wing capability). This rapid folding and unfolding stems from the microstructure of

the bat wings. Swartz et al. (27) studied in details the microstructure of bat wing membrane

skin. Figure 2.6 shows overall structure of a bat wing with reinforcing frame (shoulder bone,

five fingers and hind limb). The membrane structure is a network of fibers which are spun

orthogonally and support thin skin. This corrugated skin structure helps easy folding and

unfolding actuations. The skin structure along the vertical direction exhibits high stiffness while

22

Page 55: BaTboT: a biologically inspired flapping and morphing bat robot

2.3 Bat flight research

that along the horizontal direction has higher failure strain. The high stiffness of the wing along

the vertical direction is needed to keep the wing shape. Also, the surface of the membrane

is equipped with touch-sensitive receptors on small bumps called Merkel cells, found in most

mammals including humans, similarly found on our finger tips (66). These sensitive areas are

different in bats as each bump has a tiny hair in the center, making it even more sensitive and

allowing the bat to detect and collect information about the air flowing over its wings, thereby

providing feedback to the bat to change its shape of its wing to fly more efficiently.

2.3 Bat flight research

Evolutionary biologists are deeply interested in understanding and quantifying all aspects of

bat flight in terms of aerodynamics, navigation, behavior and so on. In (5) a framework for 3D

reconstruction and analysis of bat flight maneuvers is presented. The reconstructed model of

the bat is composed by 52 degrees of freedom, giving a great insight about how bats modulate

the articulations of the whole body and wings. Figure 2.7 shows the tracking of bat flight in

real time.

Figure 2.7: Tracking results of a bat using a 52 degree of freedom articulated model. Shown on

top are frames extracted from high speed video of a landing bat. Shown on the bottom are the

corresponding frames of the reconstructed three-dimensional wing and body kinematics. –Caption

extracted from (5)–.

Using this model in conjunction with the measured wing kinematics, (5) has shown that

surprisingly modulation of wing inertia plays the dominant role of reorienting the bat with little

or no reliance on aerodynamic forces.

Thus, bats are able to change body orientation mid-flight without relying on

aerodynamic forces and instead by changing the mass distribution of its body and

wings.

23

Page 56: BaTboT: a biologically inspired flapping and morphing bat robot

2.3 Bat flight research

Figure 2.8: Reconstruction of the wake of C.

brachyotis bat. Source: (6).

Another remarkable research regarding

the relationship of bat flight kinematics and

wake structure is presented in (6). Wind-

tunnel measurements of the bat specie

Cynopterus brachyotis have revealed the

resultant wake velocities produced by the

kinematics motions of the wings. Figure

2.8 shows that a closed loop vortex struc-

ture is dominant at relatively slow forward

flight speeds (4.3m/s), and there is evi-

dence for additional small vortex structures

shed from other appendage. Besides at-

tempting to quantify the kinematics of bat

flight and wake vorticity, the wing mem-

brane is

a material with key features that make bat flight unique among nature. In (7). in-flight wing

membrane strain measurements have been carried out. Data were collected from wind tunnel

wind-off flights of a Jamaican fruit bat, Artibeus jamaicensis, is shown in Figure 2.9.

Results have shown that strain levels are around 10% in the X direction and 3% in the Y

direction. Surface snapshots of the membrane strain-state show in the X direction (spanwise)

a consistent strain-relief effect around the ring finger during downstroke, with high-strain con-

centration areas on the membrane between the little finger and the ring finger. The estimates

of the shape and positions of two section of the hand wing during the downstroke

were evaluated and revealed that in order to achieve this strain distribution, the

membrane should have a mid downstroke camber value that is a function of the

body mass of the specimen Mb. This relationship has been defined as: wing camber

∝ M0.9b . Depending on the specimen’s size, this value typically varies from 0.07 to

0.25, as measured in (20).

Continuing with the membrane, the material of what it is made is not the whole thing.

There is biological evidence that bats have tiny hairs in the membrane that sense airflow

conditions, and there is some evidence that this sensing apparatus in bats contributes to their

flight efficiency (cf. Figure 2.10). In (8), evidence that the tactile receptors associated with

24

Page 57: BaTboT: a biologically inspired flapping and morphing bat robot

2.4 Shape Memory Alloys

Figure 2.9: Strain measurements of bat’s

wing membrane. Source: (7).

Figure 2.10: Sensory wing hair before (A)

and after (B and C) depilation. (A) Scanning

electron microscope image from a domed hair

located on the ventral trailing edge (location

is marked by a gray circle in schematic to the

Right) of Eptesicus fuscus. Caption extracted

from Source: (8).

these hairs are involved in sensorimotor flight control by providing aerodynamic feedback. It was

found that neurons in bat primary somatosensory cortex respond with directional sensitivity

to stimulation of the wing hairs with low-speed airflow. Wing hairs mostly preferred reversed

airflow, which occurs under flight conditions when the airflow separates and vortices form. This

finding suggests that the hairs act as an array of sensors to monitor flight speed

and/or airflow conditions that indicate stall .

2.4 Shape Memory Alloys

Within the need of building smaller devices with integrated features of sensing and actuation,

the field of smart materials have opened a new generation of actuation devices (43), (62). The

word ”smart” has been used to highlight the property of some materials capable of changing

its physical properties upon certain conditions. Some of these materials are based on: i) Piezo-

electricity, which react to the application of an electric field, ii) Shape Memory Effect (SME),

which possess the ability to actuate when subjected to thermal changes and recover its initial

configuration without any thermal process involved, iii) Electro Active Polymers (EAPs), either

ionic or electronic, which directly exploit Maxwell forces or the electrostrictive phenomenon to

25

Page 58: BaTboT: a biologically inspired flapping and morphing bat robot

2.4 Shape Memory Alloys

obtain mechanical energy from electrical input energy, iv) Electrorheological fluid (ERF), whose

rheological characteristics vary, depending on the external fields applied, and v) Magnetostric-

tion, either positive or negative, where magnetic domains are reoriented by means of an external

magnetic field.

This thesis involves the use of Shape Memory Alloys (SMAs) as both actuators and sensors

of wing motions. Basically, the SMAs make use of the Shape Memory Effect, exhibiting a

thermally activated martesitic transformation. The development and application of SMAs

attained the requisite momentum following the discovery of NiTi alloys in 1963 at the Naval

Ordnance Laboratory. The name nitinol refers to Nickel, Titanium, NOL (Naval Ordnance

Laboratory). The implementation of applications making use of SMAs has evolved hand-in-

hand with the development of nitinol. In the last few decades, both industry and academia have

evinced growing interest in the application of nitinol, and of SMAs in general. This is basically

because these materials are intrinsically susceptible of use both as sensors and as actuators,

which makes them suitable for use as smart actuators and for integration in smart structures.

This section provides a concise description about the use of SMAs as actuators in robotics,

by describing how the material operates and some of the applications in terms of modeling and

control. Also, state-of-the-art Micro Aerial robots that use smart materials for morphing-wing

actuation are highlighted from the literature.

2.4.1 Basic foundations

Fist Principles

In SMAs, the shape memory mechanism is based on a reversible, solid-state phase transforma-

tion between the high temperature austenite phase and the low temperature martensite phase.

This phase transition is also known as martensitic transformation. The martensitic phase trans-

formations of the alloy can be characterised by four transformation temperatures: i) As, the

austenite start temperature, ii) Af , the austenite finish temperature, iii) Ms, the martensite

start temperature, and iv) Mf , the martensite finish temperature. These transformations are

shown within the hysteresis curve of the SMA in Figure 2.11a.

When the temperature is less thanMf , the NiTi alloy consists only of the martensite phase.

As the temperature is increased beyond As, austenite begins to form in the alloy and when

the temperature exceeds Af , the alloy is primarily in the austenite phase. As the alloy cools,

martensite begins to form when the temperature drops below Ms, and when the temperature

26

Page 59: BaTboT: a biologically inspired flapping and morphing bat robot

2.4 Shape Memory Alloys

(a) (b)

(c) (d)

Figure 2.11: a) The hysteresis curve of SMA, b) stress-strain-temperature curve of SMA ex-

hibiting the one-way shape memory effect, c) stress-strain-temperature curve of SMA exhibiting

the two-way shape memory effect, d) stress-strain-temperature curve of SMA exhibiting pseudoe-

lasticity behavior. Source: (9).

reaches Mf , the alloy is again fully martensitic. As mentioned, these transitions represent the

SMA thermal hysteresis loop.

During phase transitions between martensite and austenite, most of the physical properties

of SMAs vary. These include Young’s Modulus, electrical resistance, heat capacity and thermal

conductivity (9), (67). Considerable research for modeling the microscopic and macroscopic

behavior of SMAs has been carried out in the last decades. Since the mechanical behavior is

closely related to microscopic phase transformations, stress-strain relations are not applicable

to describe the shape memory effect.

A single SMA wire describes one-way shape memory effect when the alloy shows permanent

deformation after the removal of an external force. Therefore, it can recover its original shape

upon heating. Subsequent cooling does not change the shape unless it is stressed again. In

addition, a two-way shape memory effect may occur upon cooling and without the applying of

external stress. This phenomenon can be easily observed within an antagonistic arrangement

of SMA wires acting in parallel. Figure 2.11b depicts the one-way shape memory effect. By

heating the deformed martensite past As, the austenite start temperature, austenite begins to

27

Page 60: BaTboT: a biologically inspired flapping and morphing bat robot

2.4 Shape Memory Alloys

form and the material begins to contract. Full shape recovery can be achieved by heating above

Af , where the alloy is completely in the austenite phase again. As this shape recovery only

occurs in one direction, it is referred to as the one-way shape memory effect.

On the other hand, the two-way shape memory effect (as described in Figure 2.11c) can

be defined as the reversible shape change upon thermal cycling in the temperature range of

martensitic transformations without requiring any external load. SMAs can be trained to

exhibit the two-way effect using two methods, which are spontaneous and external load-assisted

induction (68), (69). However, the shape change obtained is in practice less than the one-way

effect.

Another interesting phenomenon is the pseudoelasticity, which is the shape recovery associ-

ated with mechanical loading and unloading of SMAs at temperatures above Af . Figure 2.11d

depicts stress-strain temperature curve associated with the pseudoelasticity behavior of a 2D

crystal structure model of SMAs. There is no temperature change required for pseudoelastic

behaviour. Therefore, the strain characteristic can be described using only the stress-strain

plane (9).

Modeling

There are two classes of SMA models: microscopic and macroscopic. The former is based on

first principles models, whereas the latter is based on experimental results provided by system

identification. In this thesis, both models will be approached, the former for simulation purposes

and the latter for control design.

The microscopic model uses in this thesis has been presented by Elahinia in (56) and (70).

The model is described as a multi-dimensional thermomechanical constitutive phenomenological

model. The model is based on Tanaka’s pioneer work in (71). Details on this microscopic model

can be found in Section 4.5.2. Similar phenomenological models can be found in the literature:

Kuribayashis model based on experimentally identified relations (72), (73), which observed a

linear relationship between very small variations in the force and strain of an SMA wire. Under

constant strain, the relationship between the force and supplied voltage was also observed to be

approximately linear. The sublayer models of Hirose et al. (74), (75) introduced the two-phase

model for SMAs using a commonly method in solid mechanics to describe nonlinear stress-

strain relationships. The model considers the SMA to be composed of parallel sublayers of the

different phases with their respective mechanical properties. This is combined with a model of

transformation kinetics based on thermodynamics to form the variable sub-layer model. Laterm

28

Page 61: BaTboT: a biologically inspired flapping and morphing bat robot

2.4 Shape Memory Alloys

Madill et al., in (76) extended their work for a new SMA actuator model that is capable of

modeling minor hysteresis loops. Others, such as Liang and Rogers (77), Brinson (78) and

Elahinia (70) improved upon Tanaka’s model using different transformation kinetic equations

relating the martensite fraction to the stress and temperature.

To approach macroscopic modeling based on system identification process, this thesis uses

the procedure described in (9) which consists on small-signal response of SMA wires over a

suitable frequency range aimed at defining the relationship between input power and the output

force of the wire during contraction. As demonstrated by Yee in (9), Kuribayashi in (73), and

Grant (79), the power-to-force relationship of the AC response of a NiTi SMA wire can be

modeled similar to a first-order transfer function. Section 6.3 will detail on this issue.

Actuation and Control

Basically SMAs are used as two modes of actuation: one actuator composed of a SMA wire and

a bias spring, or two SMA wires featuring an antagonistic configuration. Herein we have used

the latter, also called, differential SMA actuator. In 1984, Honma et al., (80) demonstrated

that it is possible to control SMA actuation by electrical heating. They used open loop Pulse

Width Modulation (PWM) to operate the actuation. The first control of an antagonistic pair

of SMA wires for moving a rotational joint was presented by Kuribayashi in (72). PWM was

used in a feedback controller to control both position and force. The control scheme adjusted

the duty cycle between the two antagonistic actuators by switching the applied voltage between

the two actuators.

Currently SMA control can be divided in three categories: i) Linear control, ii) PWM,

and iii) non-linear control. Linear control is widely extended, featuring PID methodology for

the control strategy. Some of the more notable work include (81), (82), (83), (84), and more

recently, (9) and (67). On the other hand, PWM technique has been implemented by some of

the early SMA researchers including (80), (82), (72), and more recently, (85).

Non-linear control of SMAs is also widely extended. Pons et al., in (58). compared PI control

based on direct strain feedback linearisation and feedforward approach to the conventional PI

controller. It was shown that feedforward control achieved the best overshoot reduction. Other

comparisons with linear controllers include Lee and Lee [43] who investigated time delay control

on SMA actuators, and Ahn and Nguyen [1] who experimented with self-tuning fuzzy PID

controllers.

29

Page 62: BaTboT: a biologically inspired flapping and morphing bat robot

2.4 Shape Memory Alloys

In terms of improving the actuation speed of SMAs, Yee et al., (9), (67) studied different

phenomena on NiTi SMA wires, from small-signal high frequency response analysis, to force

models based on system identification, and the proper mechanisms that allow for faster and

accurate force control of an antagonistic pair of SMA actuators. These mechanisms are called:

i) anti-slack, and ii) anti-overload. The former deals with the two-way shape memory effect (62),

improving accuracy and speed, whereas the latter limits the amount of input heating power

to prevent physical damage when SMAs are overloaded. As a result, their force controller

was capable of tracking fast and accurate force references compared to other works from the

literature (86), (79), (87). Nonetheless, their control architecture requires of a high-bandwidth

force sensors capable of providing the force feed-back.

In this thesis, a control architecture similar to the one described in (67) has been imple-

mented, which makes use of both anti-slack and anti-overload mechanisms to manage the afore-

mentioned limitations in SMA actuation speed and accuracy. However, some important changes

to the architecture are proposed, since the available payload capacity of the bat robot constrains

the use of force sensors. Instead, SMA electrical resistance feed-back is used for sensing mo-

tion (88), (89). Thereby, both anti-slack and anti-overload mechanisms have been conceived for

regulating the amount of input heating power based on a Resistance-Motion (RM) relationship.

This technique will be detail on Section 6.4, while experimental results in Section 6.5 will show

proper position control performance in terms of tracking and actuation speed of the SMAs,

applied to the morphing-wing mechanism of the bat robot.

2.4.2 Advantages and drawbacks

Most relevant advantages and drawbacks of using SMA technology for actuation are highlighted

as follows:

Advantages

• Size and weight : SMAs can be directly used as linear actuators. There is no need for

additional motion components or hardware, which permits easy miniaturizations of the

actuation system. For the application at hand, BaTboT requires the minimum added

weight in order to be able to fly. Other actuation mechanisms, such a DC-motors, servos,

etc, are simply inappropriate due to their weight. Instead, SMA wires have a negligible

volume (e.g., 3× 10−9m3), allowing for extremely light wings.

30

Page 63: BaTboT: a biologically inspired flapping and morphing bat robot

2.4 Shape Memory Alloys

• High Force-to-weight ratio: SMA actuators have a large force-weight ratio of ∼ 8N/1 ×10−5Kg, using a wire with thickness of 150μm, and 0.1m long. SMAs also present large

life cycles (3× 106).

• Noise-free operation: Because SMA actuators do not require friction mechanisms such as

reduction gear, it avoids the production of dust particles, sparks and noise. These merits

make SMA actuators extremely suitable for areas such as microelectronics, biotechnology

and biomimetics applications (high bio-compatibility).

• Sensing properties: Although SMAs are mostly used for actuation, they also have sensing

capabilities. Several properties of the SMAs change as it undergoes martensite phase

transformation. Among these properties is the resistivity that decreases as the tempera-

ture of the wire increases and hence its phase transforms to austenite. A liner relationship

between electrical resistance change and SMA strain can be derived.

Drawbacks; challenges to tackle

• Slow speed : SMA actuators have generally been considered to have slow response due to

restrictions in heating and cooling, and also due to the inherent thermal hysteresis. The

common method in actuation is by electrical heating. Although applying larger electrical

currents can increase the speed, this may also overheat and damage the actuator without

monitoring. Most research so far has investigated SMA position control at generally low

tracking speeds of less than 1Hz. Rise times for step responses usually took more than 1

second.

• Fatigue: Long-term performance of the Shape Memory effect could decrease over time if

the material is expose to large external stress or overheating temperatures resulting from

large input currents.

• Low energy efficiency : The maximum theoretical efficiency of SMAs is of the order of 10%

based on the Carnot cycle, according to (90). In reality, the efficiency is often less than 1%,

since the SMA actuator can be considered a heat engine operating at low temperatures.

This means that the conversion of heat into mechanical work is very inefficient. Most of

the heat energy is lost to the environment. Hence SMA actuator applications must be

limited to areas where energy efficiency is not an issue (cf. (9)).

31

Page 64: BaTboT: a biologically inspired flapping and morphing bat robot

2.5 Morphing-wing MAVs with smart actuation

a)

b)

Figure 2.12: a) The BATMAV robot, b) Detailed arm assembly using SMAs-bases muscles.

Source: (10).

2.5 Morphing-wing MAVs with smart actuation

This Section reviews the state-of-the-art Micro Aerial Robots (MAVs) that use smart materials

for actuation, more concretely Shape Memory Alloys, although not limited to this material.

Other platforms featuring smart materials, such as Piezoelectric or polymers are also included,

which are not necessarily biologically inspired.

BATMAV

To the best of the authors’ knowledge, the only works attempting to reproduce bio-inspired

bat flight using SMAs are presented in (10) and (47). A robotic platform called BATMAV

(fully actuated by SMA wires) is described in both papers. Thereby, SMAs have been used for

two purposes: first, as muscle-like actuators that provide the flapping and morphing wingbeat

motions of the bat robot, and second, as super-elastic flexible hinges that join the wings’ bone

structure. The BATMAV prototype is shown Figure 2.12.

32

Page 65: BaTboT: a biologically inspired flapping and morphing bat robot

2.5 Morphing-wing MAVs with smart actuation

a)

b)

c)

d)

Figure 2.13: a) Bat-wing platform in the wind-tunnel, b) Nylon membrane, c) Spandex mem-

brane, d) Silicone membrane. Source: (11).

Most of the experiments in (10) were carried out with a two degree of freedom wings capable

of flapping at 3Hz. Nonetheless and despite the robot is able to achieve similar bio-inspired

trajectories, the results lack of experimental evidence relating aerodynamics measurements that

demonstrate the viability of the proposed design. Moreover, neither (10) or (47) detail how to

control the SMAs for achieving the bio-inspired motion of BATMAV’s wings.

Bat-Wing

The research presented in (11) is aimed at the study of bat wing biomechanics and aerody-

namics performance for the design of an artificial wing. The wing prototype is composed by a

basic skeletal structure based upon relative dimensions from the anatomy of a bat. Different

wing models were assembled comparing different materials for the membrane, such as: Nylon,

Spandex, and Silicone. Wind tunnel measurement have been carried out to analyze the wing

performance in terms of Lift-to-drag production. Figure 2.13 shows the Bat-wing prototype.

The results have shown that that silicone is simply a more aerodynamically-suited material.

This material is non-porous, is able bend along with a flexible skeleton to create optimal flight

shapes, and appears to be able to withstand contact with winds up to 20mph to create lift

forces.

Another interesting research from the same authors (48) has also studied key features of bat

flight in order to understand how certain aspects of the wings might improve on the design for

micro air vehicles. The goal of the study is not to mimic natural bat flight, but to understand

33

Page 66: BaTboT: a biologically inspired flapping and morphing bat robot

2.5 Morphing-wing MAVs with smart actuation

veins

moldedpolymer

carbonfiber

piezo-motor

bi-morphingtransmission

(thorax)

power electronics

Figure 2.14: Harvard RoboBee, actuated by PZT. Source: (12).

how certain aspects of bat flight apply to the engineering problem of wing design for micro air

vehicles. Aspects such as morphing, cambering, and twisting are measured and quantified in

terms of lift production and drag reduction/

Robobee

Harvard’s scientist in (12) present an innovative insect-scale robotic thorax designs capable of

producing asymmetric wing kinematics similar to those observed in nature and utilized by flies

and other two-winged insects to maneuver. Inspired by the thoracic mechanics of such insects,

a Piezoelectric (PZT) actuator has been fabricated as a bending bimorph cantilever actuator.

The transmission maps the approximately linear motions of the actuators into the flapping

motion of the wings. The transmission consists of links and joints with geometries designed to

maximize the product of stroke amplitude and first resonant frequency, given known actuator

and airfoil properties. The insect-like robot, shown in Figure 2.14 is capable of flapping at 90Hz.

Actuators are created by laminating two piezoelectric plates (PZT-5H from Piezo Systems,

Inc.) to a carbon fiber spacer and electrode layer. The airframe and transmission are created

by layering a 7.5μm polyimide film (Chemplex Industries) with carbon fiber face sheets.

Morphing trailing edge

A trailing edge being developed by Wildschek et al., in (13) splits the trailing edge into seamless

top and bottom control surfaces that are morphed independently. The inner structure consists

of hollow segmented parts separated by bars and hinges that allow for both compliance and

34

Page 67: BaTboT: a biologically inspired flapping and morphing bat robot

2.5 Morphing-wing MAVs with smart actuation

Figure 2.15: Inner structure and possible control modes of a multi-functional trailing edge.

Source: (13).

aerodynamic loading. The model has been built out of carbon reinforced plastic that is mor-

phed through the use of an all- electric actuation system. The model is capable of multiple

actuation modes (pitch, roll, high lift, etc). Within the airfoil, SMAs have been used in a spring

configuration for providing the wing actuation. The wing prototype is shown in Figure 2.15

and could be used for improving fixed wing current aircrafts’ design.

Morphing-wing camber MAV

Wilkie et al., in (14) have developed a MAV capable of varying the camber of its wings. Varying

the camber in a wing can have beneficial properties for the control of an aircraft such as during

take-off and landing when the lift distribution along a wing is required to dramatically change.

The MAV is depicted in Figure 2.16, which consists of a macro-fiber composite (MFC); a

flexible film with a layer of unidirectional piezoceramic fibers sandwiched between layers of

copper electrodes and acrylic/Kapton. Wind tunnel tests revealed improved drag characteristics

due to having a continuous wing surface as opposed to discrete articulated control surfaces, a

characteristic shared by many morphing structures. Asymmetric and symmetric actuation of

the MFC patches provides roll and pitch moments of around 0.06Nm.

Other morphing-wing MAV concepts

In general, most of the morphing-wing MAV concepts out there relies on standard motor ac-

tuation technology. Some of the best concepts are provided by FESTO smart-bird (cf. Figure

2.17a) (15), where its wings not only beat up and down, but also twist at specific angles. This

is made possible by an active articulated torsional drive unit, which in combination with a

complex control system attains an unprecedented level of efficiency in flight operation. Festo

35

Page 68: BaTboT: a biologically inspired flapping and morphing bat robot

2.5 Morphing-wing MAVs with smart actuation

Figure 2.16: Layout of an MFC actuation device for MAV morphing-wing camber. Source: (14).

Figure 2.17: TOP (left to right) Smart-bird by FESTO, inspired by the Seagles (15), Combat by

University of Michigan, inspired by bat’s navigation system (16), DARPA nano-air hummingbird

(http://www.darpa.mil). BELOW (left to right) The Gull Wing by (17), MFX-1 by NextGen

Aeronautics, inspired by the batwing internal structure (18), Prototype of Entomoter MAV (19).

has thus succeeded for the first time in creating an energy-efficient technical adaptation of this

model from nature.

University of Michigan researchers have came out with a bat-llike MAV concept, shown in

Figure 2.17b (16). The project is funded by DARPA, aimed at the development of a robotic

bat with the ability to navigate at night, using low-power miniaturized radar and a very sen-

sitive navigation system. The robotic bat will also have the ability to navigate at night, using

low-power miniaturized radar and a very sensitive navigation system. Its lithium battery will

recharge using solar energy, wind, and vibrations, and the bat will communicate with the troops

using radio signals. Bats have a highly-attuned echolocation sense providing high-resolution

navigation and sensing ability even in the dark, just as our sensor must be able to do. Echolo-

cation allows bats to navigate by emitting sounds and detecting the echoes. The robot’s body

concept is designed to be about six inches long and to weigh about a quarter of a pound. Its

expected energy consumption will be 1W .

Lind et al., in (17) have developed and built a series of UAV models that incorporate

36

Page 69: BaTboT: a biologically inspired flapping and morphing bat robot

2.6 Remarks

a folding wing concept called a Gull Wing (cf. Figure 2.17c) The wings contain a telescoping

spar connected to a hinged spar that enables the folding motion. Another MAV that takes some

inspiration by bats’ wings design is presented in (18) (cf. Figure 2.17d). The MAV is called

the NextGen batwing, developed by NextGen Aeronautics. The MAV successfully underwent

a 40% planform area morphing, 30% wingspan morphing, and 20o sweep angle morphing in

mid-flight. Further studies have been made on optimization of the batwings cell structure and

actuator placement, as well as developing control laws to efficiently control the wings morphing.

2.6 Remarks

Bats have enormous advantages in both inertial and aerodynamics compared to other flying

animals. Their high body-to-wing has ratio and high wing dexterity allow bats to perform aggres-

sive maneuvers modulating solely wing inertial. Also, changing the wing profile improves of the

generation of lift forces and reduction of drag during the wingstroke. These factors might be the

key for developing micro aerial systems that attempt to mimmic efficient flapping/morphing

wing flight. From the biological review presented in this chapter one can immediately note why

the bat apparatus is worthy to mimic using BaTboT:

• Low wingbeat frequencies : as shown in Figure 2.2d, bats can produce enough lifting forces

by flapping their wings in the range of ∼ 10Hz. To achieve that, bats have larger

wingspans (Figure 2.2a) and low body mass. This combination is essential into the proper

production of lift. Note e.g., in Figure 2.2d, that Hummingbirds in the same body mass

scale compared to bats, require to flap their wings at least five times faster than bats. In

terms of design, the scale values for fabricating a robot with similar wingspan and body

mass are perfectly achievable with BaTboT.

• Morphing wings and cambering : bat wings are unique in nature. They contain powerful

muscles that provide the same dexterity and mobility (degrees of freedom) than the human

arms. It means they can fold, expand and camber their wings in such a unique way.

Furthermore, the wing membrane also contains embedded tiny muscles and veins that

contribute in controlling the tension of the membrane during flight (cf. Figure 2.6).

• Sensing : as mammals, bats are actually able to see very well. By combining great sight

and echolocation, they can maneuver in narrow spaces, during daylight or night. In

addition, bat’s wings have tiny hairs over the membrane’s surface that can sense the

airflow and adjust wing morphology to improve on aerodynamics.

37

Page 70: BaTboT: a biologically inspired flapping and morphing bat robot

2.6 Remarks

• Efficient flight : By combining both morphology and sensing capabilities, bats are the best

flying animals to maneuver at low speeds and altitudes than any biological mechanism

in nature. They can hover like a hummingbird, maneuver like birds, or change direction

abruptly like insects. On top of that, bats use 35 percent less energy by reducing aero-

dynamic drag, having the lowest power-to-mass ratio (W/g) of any flying vertebrate in

nature (cf. Figure 2.3d).

In terms of actuation, Shape Memory Alloys (SMAs) enable the fabrication of lighter wins

with muscle-like actuation but some challenges should be addressed. Section 2.4.2 highlighted

the advantages and drawbacks of this material. This thesis will present feasible solutions to

minimize the effects of SMA limitations and it will give an insight into the performance of the

material acting as actuators. The goal is not only to evaluate the use of this actuation technology

for the application at hand but also on providing a formal quantification of performance that

would allow others to drive this technology forward.

By reviewing the state of the art in Section 2.5, one can note the lack of biologically-inspired

robots that explore alternative actuation mechanisms more likely to those found in nature. The

field of bio-inspired MAVs that use smart materials for actuation is still in an early stage. Most

of the works have investigated how to fabricate efficient wings models, but few have achieved

to develop a complete bio-MAV platform capable of sustained flight. This thesis embarks into

this potential field by presenting the first highly articulated bat-like MAV that can maneuver by

means of changing wing morphology and also it takes advantage of the improvements in flight

efficiency that wing modulation enables.

38

Page 71: BaTboT: a biologically inspired flapping and morphing bat robot

3

From bats to BaTboT:

Mimicking biology

3.1 General overview

This chapter presents insights of in-vivo bat flight. BaTboT’s morphology and biomechanics

are based on the bat specie Rousettus aegyptiacus physiology. This section describes why the

selection of this specie to be mimicked with BaTboT. This selection has been based on criteria

regarding: i) wing morphology (i.e., wingspan, aspect ratio, body and wing mass, etc), ii) wing

kinematics (wing joint trajectories), and iii) wing aerodynamics (wingbeat frequency, angle of

attack, lift production, etc). The biological data is analyzed from in-vivo biological experiments

of several species carried out in (22), (23), (24), (20).

3.2 Review on biological flight performance data

Body size governs almost every aspect of animal biology. In bats, body size and mass, wing

mass and wingspan determine most of the kinematics, dynamics and aerodynamics performance

of the animal. Hence, it is reasonable to think that a proper choice of the specimen to mimic

should be based on the analysis of these parameters. (20) presents a rigorous study that

compares the wing kinematics of 27 bats representing six pteropodid species ranging more than

40 times in body mass (0.0278−1.152kg). This section briefly summarizes the results presented

in (20), relating and highlighting performance issues mostly in terms of wing kinematics and

aerodynamics.

39

Page 72: BaTboT: a biologically inspired flapping and morphing bat robot

3.2 Review on biological flight performance data

1

2

3

zo

xo{o}

wingtip

wristc d

ef

g

h

kjilm

n

qp

ob a

Figure 3.1: Right lateral view of the bat with respect to the inertial coordinate system {o}.Green dots are the path of the wrist joint whereas red dots are the path of the wingtip over a

wingbeat cycle. The position and posture of the right wing are shown at three time points in the

wingbeat cycle. Source: (20).

The biological experiments consigned in (20) consisted in recording the flight kinematics of

27 animals from six species shown in Table 3.1. The positions of 17 anatomical markers placed

on the individual’s wings were digitized via high-speed cameras. Five flights of the individuals

within the wind-tunnel were used for analyses. The bats were recorded performing forward flight

and describing the wing-trajectory profile shown in Figure 3.1.

Table 3.1: Description of the 27 individuals used in the study. Source: (20).

Species name Abbreviation (Color code) Body mass of specimens [kg]

Cynopterus brachyotis Cb(Purple) 0.028, 0.031, 0.035, 0.035, 0.040

Rousettus aegyptiacus Ra(Blue) 0.112, 0.132, 0.159

Pteropus pumilus Miller Pp(Green) 0.178, 0.178, 0.180, 0.204, 0.212

Eidolon helvum Eh(Yellow) 0.254, 0.266, 0.278, 0.326, 0.332

Pteropus hypomelanus Ph(Orange) 0.454, 0.464, 0.468, 0.490, 0.526

Pteropus vampyrus Pv(Red) 1.020, 1.052, 1.090, 1.152

3.2.1 Measurements of wing morphological parameters

The following parameters were calculated regarding how the specimens from Table 3.1 changed

their wings during the trials (cf. Section 2.2 for detailed definition of morphological parameters):

40

Page 73: BaTboT: a biologically inspired flapping and morphing bat robot

3.2 Review on biological flight performance data

Figure 3.2: (A) Maximum wingspan, (B) minimum wingspan, (C) wing chord, (D) maximum

wing area, (E) wing loading, and (F) aspect ratio. Circles represent medians for each species and

the black arrow points to the specimen under analysis. Source: (20).

• Maximum and minimum wingspan (bmax, bmin): bmax is two times the maximum distance

of the wingtip marker from the mid-sagittal plane yo = 0 (measured during downstroke).

bmin is the opposite (measured during upstroke).

• Maximum wing chord (cmax): the longest two-dimensional distance between the wrist

and the tip of digit V (e and n in Figure 3.1).

• Maximum wing area (Smax): the left wing was divided into 18 triangular surfaces (see

Figure 3.1). The areas of those triangles were summed, then multiplied by two, to arrive

at total wing area (S).

• Wing loading (Qs) and Aspect ratio (AR): defined as Qs = mtgS−1max, being mt the bat’s

mass, g is the gravity, and Smax the maximum wing area. AR = b2maxS−1max.

Figure 3.2 compares how morphological parameters among the specimens are affected de-

pending on the body mass mt. This study allows for the quantification of scaling factors that

determine relationship between morphological parameters and body+wing mass. Table 3.2

details these values.

3.2.2 Measurements of kinematics parameters

Similarly, kinematics parameters are shown in Figure 3.3. The horizontal flight speed scales

as: V ∝ m0.005t . The horizontal velocities of bats (4.98 ± 0.09ms−1) were much greater than

41

Page 74: BaTboT: a biologically inspired flapping and morphing bat robot

3.2 Review on biological flight performance data

Table 3.2: Scaling factors for wing morphological parameters as a function of body+wing mass

mt (cf. Figure 3.2). Source: (20).

Parameter Scaling factor

Minimum wingspan (bmin) bmax ∝ m0.423t

Maximum wingspan (bmax) bmin ∝ m0.366t

Maximum wing chord (cmax) cmax ∝ m0.357t

Maximum wing area (Smax) Smax ∝ m1.32t

Wing loading (Qs) Qs ∝ m0.233t

Aspect ratio (AR) AR ∝ m0.072t

Figure 3.3: (A) Flight speed, (B) Horizontal accelerations, (C) Vertical accelerations. Circles

represent medians for each species and the black arrow points to the specimen under analysis.

Source: (20).

vertical velocities (0.12± 0.03ms−1), so flight paths were close to horizontal plane (1.36± 0.36o

above horizontal).

3.2.3 Measurements of Aerodynamics parameters

Figure 3.4 compares the most relevant parameters that indirectly affects the aerodynamics of

the bats. Scaling factors are detailed in Table 3.3.

Table 3.3: Scaling factors for wing aerodynamics parameters as a function of body’s mass mt

(cf. Figure 3.4). Source: (20).

Parameter Scaling factor

Minimum wingbeat period (T ) T ∝ m0.180t

Downstroke duration (Tdown) Tdown ∝ m0.213t

Downstroke ratio (τdown) τdown ∝ m0.036t

Maximum wing stroke amplitude (φs) φs ∝ log(mt)(−3.058)

Strouhal number (St) St ∝ m−0.088t

Minimum angle of attack (α) α ∝ log(mt)(−7.738)

Wing camber at maximum span camber ∝ m0.9t

Minimum lift coefficient (CL) CL ∝ m0.170t

42

Page 75: BaTboT: a biologically inspired flapping and morphing bat robot

3.3 Choice of species

Figure 3.4: Wingbeat period (A) scaled lower than expected under isometry. Downstroke dura-

tion (B), Downstroke ratio (C), stroke amplitude (D), stroke plane angle (E) and Strouhal number

(F) did not change significantly with body mass. Angle of attack increased with body size (G) as

a result of a change in α1 (H), but not from a change in α2 (I). Wing camber (J) did not change

with body size, but coefficient of lift (K) did. Circles represent medians for each species and the

black arrow points to the specimen under analysis. Source: (20).

3.3 Choice of species

From the starting point of this thesis, the design process of BatBot was visualized and thought

within the range of bio-inspired Micro Air Vehicles (MAVs). Unfortunately there is no standards

or normative that strictly classify the most relevant properties of bio-MAVs. However, the

literature commonly use wingspan and mass properties as criteria for classifying different types

43

Page 76: BaTboT: a biologically inspired flapping and morphing bat robot

3.3 Choice of species

Birds

UAVs

Aircraft

Bio-MAV design scale

Insects

mas

s (K

g)

wingspan (m)10.10.01 10 100

1

0.1

0.001

100

10000

Bats

BaTboT

Figure 3.5: Bioinspired MAV classification

depending on wingspan and mass. Source:

the author.

Figure 3.6: Wing physiology. 17

markers are placed on: anterior

and posterior sternum (a and b, re-

spectively), shoulder (c), elbow (d),

wrist (e), the metacarpophalangeal

and interphalangeal joints and tips

of digits III (f, g, h), IV (i, j, k), and

V (l, m, n), the hip (o), knee (p),

and foot (q). Source: (21).

of Unmanned Air Vehicles (UAVs). Figure 3.5 details this classification.

In nature, bat’s wingspan varies from 0.15m (Craseonycteris thonglongyai micro bat) to

1.82m (Pteropus vampyrus mega bat). Likewise, the mass of each specimen varies from 8g to

1kg respectively. From the specimens considered in (20) whose study was briefly summarized

in the previous section, wingspan varies from 0.3m (Cynopterus brachyotis) to 1.1m (Pteropus

vampyrus Linnaeus), whereas mass varies from 0.028Kg to 1.15Kg respectively.

The selection of the specimen to mimic is based on four main criteria: i) efficiency

in lift production at ii) lowest flapping frequency with iii) about half-meter wingspan and iv)

minimum weight.

The Rousettus aegyptiacus also known as the Egyptian fruit bat fulfills with these criteria

(cf. blue-color dot in Figures 3.2, 3.3, 3.4). Another specimes with similar morphology and flight

performance: Pteropus poliocephalu and Cynopterus brachyotis are also taken into account into

the design process of BaTboT. The Rousettus aegyptiacus has an average maximum wingspan of

∼ 0.5m when the wings are fully extended (cf. Figure 3.2A) and a minimum wingspan of ∼ 0.2m

when the wings are folded (cf. Figure 3.2B). Wingspan size allows for a feasible fabrication of

BatBoT’s biomechanics using standard Fused Deposition Modeling (FDM) rapid-prototyping

in 3D. Also, this specimen has an average mass of ∼ 125g (cf. Table 3.1) which allows for a

feasible weight to obtain with standard Acrylonitrile Butadiene Styrene (ABS) material.

Besides morphology, the most important characteristic of this specimen is related to its

lift production. Normally, large bats have higher coefficients of lift than small bats. This is

obviously due to large wingspan and area. However, the Rousettus aegyptiacus is capable of

44

Page 77: BaTboT: a biologically inspired flapping and morphing bat robot

3.3 Choice of species

shoulder

wrist

MCPIIIwingtip

1 2 3 4 5 6 7 8 9 10 11 12 13 14

ep

0.049m 0.057

B = 0.197m (extended, downstroke)

0.090m

0.014m

WristIII

IV

V

(a) (b)

Wingtip

a

b

c

d

ef

g MCPIII

15deg

~15deg

f =0.032

h=0.014

g=0.012e=0.045

35.4deg63.3degreference

line

y

x

reference line

III

IV

V

wing joints

Figure 3.7: Detailed parameters that describe wing segment morphology. a) wing segment

subdivision, b) detailed configuration of the wrist joint and attached digits when the wing is fully

extended. It shows the angles between digits that maintain proper wing membrane tension during

downstroke. Source: the author.

efficiently produce lift forces at low wingbeat frequencies even with a mid-size wingspan. Note

in Figure 3.4K, the lift coefficient raises up to ∼ CL = 1.15 with the bat flapping at f = 6.6Hz

(cf. Figure 3.4A, wingbeat period T = 0.15s). Comparing the Rousettus aegyptiacus against

the Cynopterus brachyotis, note the former doubles CL by flapping about 50% slower compared

to the latter.

In the following, key aspects of the specimen morphology are determined as an useful frame-

work for BaTboT modeling and design.

3.3.1 Wing morphology

Key components of wing physiology are shown in Figure 3.6. In (22), 17 markers placed

along the wing allowed the quantification of wing size and proportions. Using these markers

the wing planform is divided into segments, allowing for the morphological description of the

wing using three geometrical parameters: i) wing chord, ii) leading edge position (b), and iii)

position of the center of mass (c). Figure 3.7 details these segments with measurements for the

specimen at hand. Also, Table 3.4 shows the values of geometrical parameters that describe

wing morphology.

Measurements of leading edge position (b) and the position of the center of mass (c) are

chordwise relative to a reference line through the left and right glenohumeral joint; negative

values are below the reference line. Figure 3.7a shows the detailed wing segments, and the

parameters that completely describe the size and proportions of the wing. Five key-dots (red

45

Page 78: BaTboT: a biologically inspired flapping and morphing bat robot

3.3 Choice of species

Table 3.4: Detailed wing segment geometry at downstroke (segments according to Figure 3.7a.)

Segment Chord a[m] Leading-edge position b[m] CM position c[m]

1 0.11 -0.035 0.004

2 0.107 -0.0347 0.0035

3 0.098 -0.0294 0.005

4 0.087 -0.0081 0.093

5 0.082 0.0086 0.014

6 0.083 0.0148 0.0198

7 0.086 0.032 0.03

8 0.09 0.036 0.032

9 0.092 0.0276 0.032

10 0.079 0.030 0.038

11 0.061 0.032 0.044

12 0.051 0.028 0.037

13 0.028 0.029 0.032

14 0.01 0.033 0.029

colored), corresponding to segments: #1, 4, 7, 11, and 14, are placed at the i) shoulder, ii)

elbow, iii) wrist, iv) MCP-III medium point, and iv) wingtip. The shoulder joint is connected

with the elbow joint through the humerus bone and the elbow is connected with the wrist

joint through the radius bone. Digits III, IV and V are connected to the wrist. The wrist

joint plays an important role into the kinematic of bat flight. Bats have very complex wrists,

similar to human hands. It allows for the rotation of digits in three-dimensions. Attempting

to mimic this complexity using an artificial counterpart is unfeasible mainly because the added

weight of required actuators to drive each digit. To simplify on this, only planar motion of

the digits is considered. This allows the digits to open and close by rotating about the gravity

axis aimed at maintaining the proper wing membrane tension during the downtroke. Figure

3.7b details the wrist and digits configuration. When the wings are fully extended, digits must

be completely open to keep the membrane with the largest tension. This improves on lift

production. To this purpose Figure 3.7b shows the maximum angles that separate one digit

from another (angles measured with respect to the joint frame of the wrist and the reference

line). The angles are calculated from key length-proportions defined at each segments. These

proportions are h = 0.014m which corresponds to the distance of each segment subdivision,

f = 0.032m which is the leading-edge position measured at segment #11 (parameter b in Table

3.4) and g = 0.012m which corresponds to the distance g = c− b at segment #11.

3.3.2 Biological-based framework for modeling and design

Concretely, Table 3.5 shows the most relevant geometrical parameters to be used for the mod-

eling and design of the robot. Also, Table 3.6 provides quantification about how morphology,

46

Page 79: BaTboT: a biologically inspired flapping and morphing bat robot

3.4 Remarks

Table 3.5: Key bio-inspired geometrical parameters for modeling and design.

Parameter [unit] Value

Total mass mt [g] 125

Extended wing length B [m] 0.195

Body width lm [m] 0.07

Extended wing span S = lm + 2B [m] 0.53

Extended wing area Ab [m2] 0.05

Humerus length lh [m] 0.055

Humerus average diameter 2rh [m] 0.0055

Radius length lr [m] 0.070

Radius average diameter 2rr [m] 0.0042

Plagiopatagium skin thickness [m] 0.0001

Table 3.6: Biological-based framework for modeling and design. mt = 0.125Kg

Parameters relation-value

mass/wing-length ∼ 0.35g/cm

Morphological Minimum wingspan m0.423t = 0.41m

Maximum wingspan m0.366t = 0.46m

Wing area m1.32t = 0.064m2

Minimum wingbeat period m0.18t = 0.68s

Kinematics Maximum wingstroke amplitude log(mt)(−3.05) = 157.8o

Minimum angle of attack log(mt)(−7.738) = 6.98o

Minimum lift coefficient m0.170t = 0.7

Aerodynamics Maximum wing camber1 m0.9t = 0.15

1 Please refer to (20) for a detailed calculation of the wing camber parameter.

kinematics and aerodynamics parameters should be defined taking into account the body+wing

mass of the robot mt. Both tables provide the biological-based framework for modeling

and design. Modeling is covered in the following Chapter 4.

3.4 Remarks

This chapter has allowed the understanding of biological parameters that directly affect bat

flight and provides the foundations and criteria for robot design. Analyses of biological ex-

periments described in (20) allowed for a complete definition of a set of key issues to consider

during the designing process of BaTboT. These issues show how morphology, kinematics and

aerodynamics can be related to each other into a bio-inspired designing framework, cf. Table

3.6 . The following chapter introduces the mathematical formulation for kinematics, dynamics,

aerodynamics and wing-actuation using SMA-like muscles.

47

Page 80: BaTboT: a biologically inspired flapping and morphing bat robot

4

BaTboT modeling

4.1 General overview

This chapter presents the modeling of the most important components involved within the de-

sign process of BaTboT: i) kinematics, ii) dynamics, iii) aerodynamics and iv) SMA for wing

muscle-like actuation.

The content of the following sections is summarized as follows:

• Kinematics: it defines the morphological parameters of BaTboT, empathizing into the

topology of the highly-articulated wings. This topology consists on two serial chains

(each wing) symmetrically connected to a floating base system (the body). Here, methods

for describing bio-inspired joint trajectories that determine the motion of the wings are

presented. The wing trajectories are aimed at producing forward and turning flight. This

requires an inertial model that will be presented in the dynamics section.

• Dynamics: it presents the inertial model. Newton Euler formalism is used to describe

dynamics Equations of Motion (EoM) of the entire bat-robot. Spatial algebra is used for

the formulation of the EoM. Spatial vectors (55) are those which the linear and angular

aspects of rigid-body motion are combined into a unified set of quantities and equations.

The notation to refer a spatial operator is X ∈ �6, being X any 6D vector. The advan-

tages of using spatial formulation is noted when writing computer code algorithms, which

makes the codes easier to read, write and debug. Using the inertial model allows the

quantification of inertial effects (forces) on thrust production and robot maneuverability.

48

Page 81: BaTboT: a biologically inspired flapping and morphing bat robot

4.2 Kinematics model

• Aerodynamics: It presents experimental quantification of lift and drag forces. It also

determines how to account for aerodynamics effects by incorporating aero terms into the

inertial model. It shows how to obtain the net forces of the system by considering both

inertial and aerodynamics contribution.

• SMA wing actuation : it defines a phenomenological model based on thermo-mechanical

equations that describe one-way shape memory effect behavior. This model is used for

the quantification of SMA performance in terms of actuation speed, output torque and

fatigue. This quantification also allows for defining the limits of SMA actuation for the

application at hand. The phenomenological model has been adapted from Elahinia’s

works in (56), (70).

Using the SimMechanics toolbox of ©Matlab, kinematics, dynamics and SMA actuation

models are integrated into a single module that represents the BaTboT’s plant. The advantage

of using SimMechanics relies on the possibility of importing all the mechanical properties from

the SolidWorks CAD model of the robot1. Simulations will be carried out aimed at determining:

• Torque requirements for wing actuation : the required torques to actuate the wing

joints are determined by the inertial model (solution of the inverse dynamics problem).

It allows for the characterization of actuators.

• Maneuverability : requirements for forward and turning flight are determined by the

inertial model. It quantifies and analyzes the influence of wing inertia on the production

of body accelerations.

• SMA limitations: the maximum allowed input electrical current that achieves the

fastest contraction and extension of the SMA actuators is determined by the phenomeno-

logical model. It allows to explore the limits to safety overload the response of the SMAs.

4.2 Kinematics model

In this section the kinematics framework of BaTboT is formulated. This framework is described

by three issues:

1. Topology. It describes the system as two chains of rigid bodies serially connected (each

wing) that are symmetrically joined to a base (body). The body is assumed to be a

1Further details about the CAD design can be found in Chapter 4.

49

Page 82: BaTboT: a biologically inspired flapping and morphing bat robot

4.2 Kinematics model

free-floating base that moves in �6. Wing frames denoted as {i} being i the wing joint

frame are placed for each degree of freedom of the wing joints.

2. Wing and body kinematics. It uses the geometrical parameters described in Table 3.5

to represent the morphology of the wings and the body. Modified Denavit-Hartenberg

(DH) convention is used to place wing frames of references, whereas Euler angles are

used to describe how the body frame rotates by following aerodynamic conventions. In

addition, spatial operators for rotation and translation based on 3x3 basic rotation matrix

are formulated. This allows for the solution of the forward kinematics problem and also

for the proper propagation of physical quantities from the wings to the body.

3. Wing trajectories. Joint trajectories for each wing are denoted by the term qi, being

the subscript i the wing joint frame. Joint trajectories have been directly extracted from

biological experiments carried out in (20), (22), (23), (24). Depending on wing modulation

joint trajectories allow for the generation of forward and turning flight. Here, simulations

to achieve both maneuvers are presented and discussed

4.2.1 Topology

Bat morphology is dimensionally complex due to the highly articulated wings. Bats have dozens

of wing joints and movement is influenced by the flexibility of the bone elements, the orientation-

dependent compliance of the membrane, their interactions with the surrounding fluid, and by

movements of the numerous tendons and muscles within the membrane. Attempting to mimic

part of that complexity is a challenge that requires identification of the most relevant joints and

its role in providing proper wing modulation. Despite bat wings are highly articulated, only six

joints are key for partially changing the wing shape. This makes the wing membrane skeletally

maneuverable by a jointed i) legs, ii) shoulder, iii) elbow, iv) wrist, and v) five fingers (digits)

connected to the wrist. Figure 4.1 shows wing and body topology.

Under this topology frames {1}-{6}R and {1}-{6}L represent the right and left wing respec-

tively, frames {b} and {0} are the body frames and frame {o} is the inertial coordinate system.

The following section details the convention to place these frames within the topology of the

robot.

4.2.2 Wing and body kinematics

Kinematics frames are described in Figure 4.2. Morphological and geometrical parameters are

defined in Tables 3.5 and 4.1.

50

Page 83: BaTboT: a biologically inspired flapping and morphing bat robot

4.2 Kinematics model

{b,0}{1,2}L

{3}L

{4,5,6}L

wingTip wingTip{1,2}R

{3}R

{4,5,6}R

{7}L {7}R

6-dof floating body {o}

zo

yoxo

body-frame {b} - 6dof floating body

frames {1},{2} - 2dof shoulder jointframe {3} - elbow jointframe {4,5,6} - 3dof wrist joint

frame {0} - base

{o} - Inertial frame

Figure 4.1: Topology. The robot has an overall of 14-DoF (not counting the 6-DoF of the floating

body). Each wing has 6-DoF and each leg 1-DoF. Source: the author.

Body-frame

The body frame {b} has xb pointing cranially along the body axis, yb pointing laterally toward

the right wing, and zb points downward and lies in the plane of symmetry of the body. The

rotation of a rigid body in space can be parameterized using several methods: Euler angles,

Quaternions, Tait-Bryan angles, etc. The most extensively used method in aerospace engi-

neering is the Euler angles, which consist in a mathematical representation of three successive

rotations about three angles: roll, pitch and yaw. Thereby the rotation of {b} with respect to

the inertial frame {o} is represented by the Euler angles: roll (φ), pitch (θ), and yaw (ψ) follow-

ing aerodynamic conventions (54). In the inertial frame {o}, xo and yo describe the horizontal

plane and +zo points in the direction of gravity. On the other hand, the base frame {0} is a

rotated body frame {b} that allows the axis x0 to point laterally toward the right wing. This

rotated body frame is aimed at applying Denavit-Hartenberg convention (53) to place wing

frames.

51

Page 84: BaTboT: a biologically inspired flapping and morphing bat robot

4.2 Kinematics model

q3lh

lrlm/2

lwt

q4,5,6

q2

q1

z1

z2

z3

x4,5,6x3

x1

wingtip

y1

y3

zb

y2x2

xb

yb

zoxo

yo {o}

y4,5,6

z4,5,6z0

x0

y0

q3

q4,5,6q2

q7

xb

yb θ

φ

Figure 4.2: Detailed description of wing kinematics frames based on Denavit-Hartenberg (DH)

convention, qi corresponds to the rotation angle from axis xi−1 to xi measured about zi. The

subscript i indicates the frame of reference (i = 1..6). The inset is a top view of the right wing

showing planar angles.

Wing-frames

As mentioned, each wing, i.e., frames from {1} to {6} is treated as a serial chain of rigid bodies

connected to a base frame {0}. Frames {0} to {6} have been placed following modified Denavit-

Hartenberg (DH) convention (53). Denavit-Hartenberg parameters consists on four geometrical

parameters that specify the position of one coordinate frame relative to another. DH coordinate

frames are placed according to the following rules:

1. Axes z0 and zn+1 are aligned with the zb axes of the base.

2. Axes z1 to zn are aligned with the n joint axes such that zi is aligned with the axis of

joint i.

3. Axis xi is the common perpendicular between zi and zi+1, directed from zi to zi+1.

Having placed the coordinate frames, their relative locations are described by the following

DH parameters. Table 4.1 shows the numerical values of DH parameters whereas Figure 4.2

depicts how these geometrical parameters allows for the kinematics representation of the wing

structure. In the following the subscript i represents the joint frame of wing.

• αi, is the angle from zi−1 to zi measured about xi−1. For the application at hand, it

represents the body/bone twist.

• ai, is the distance from zi−1 to zi measured along xi−1. It represents the body/bone

length.

52

Page 85: BaTboT: a biologically inspired flapping and morphing bat robot

4.2 Kinematics model

Table 4.1: Modified Denavit-Hartenberg parameters per wing.

Joint Frame α a q d Constraint1

Shoulder2 1 π/2 +AoA lm/2 = 0.035m q1 0 120o

Shoulder 2 −π/2 0 q2 0 80o

Elbow 3 0 lh = 0.055m q3 0 60o

Wrist+digits 4,5,6 0 lr = 0.070m q4,5,6 0 45, 30, 18o

1 Absolute rotation range.2 AoA = angle of attack.

• qi, is the angle from xi−1 to xi measured about zi. It represents the body/bone rotation.

• di, is the distance from xi−1 to xi measured along zi. It represents the body/bone offset.

In Figure 4.2, axes z1 to z6 are aligned with the six joint axes such that zi is aligned with

the axis of joint frame {i}. Axis xi is the common perpendicular between zi and zi+1, directed

from zi to zi+1. The joint angles qi are defined from axis xi−1 to xi measured about zi. The

shoulder joint of the robot is composed by two degrees of freedom: q1, q2. The former angle

allows for the primary flapping motion (rotation about axis z1) whereas the latter allows the

wings to rotate about the axis z2 (cf. insert from Figure 4.2). The elbow joint has one degree

of freedom: q3, which allows the wings to contract or extend in sync with the flapping motion.

The wrist joint has three degrees of freedom: q4, q5, q6. Each angle allows for the rotation of

digits MCP-III, IV and V about axes z4, z5, z6 respectively. Both elbow and wrist joints provide

the morphing-wing capability to the robot.

Forward kinematics

Kinematics transformations that relate two consecutive frames of the wing are given by 4x4

homogeneous transformation matrix (Ti+1,i).

Ti+1,i =

[ri+1,i pi,i+1

0 1

]=

⎡⎢⎢⎣

cos qi − cosαi sin qi sinαi sin qi ai cos qisin qi cosαi cos qi − sinαi cos qi ai sin qi0 sinαi cosαi di0 0 0 1

⎤⎥⎥⎦ (4.1)

In Eq.(4.1) ri+1,i ∈ �3x3 is the basic rotation matrix that relates frames {i+1} onto frame

{i} and pi,i+1 ∈ �3 is the position vector that joints the frame i with i + 1. The terms αi,

53

Page 86: BaTboT: a biologically inspired flapping and morphing bat robot

4.2 Kinematics model

ai, di, qi, correspond to the geometrical DH parameters of the body i. Table 4.1 details these

parameters. The forward kinematics problem can be solved as 1:

Tn,0 =n∏i=0

Ti+1,i (4.2)

In this thesis the inertial model of the robot (cf. section 4.4) will be formulated by expressing

dynamics terms using six-dimensional (6D) vectors, which leads to the development of efficient

and portable code of the algorithms (55). For this reason kinematics quantities also must

be formulated using 6D operators. This approach is defined as spatial kinematics. Spatial

kinematics formulation is convenient for the application at hand since BaTboT has a floating

body that moves in �6 and the goal is to express how both angular and linear components of

physical quantities affect into the generation of spatial accelerations and forces of the body. To

this purpose spatial kinematics using 6D operators for rotation Ri+1,i and translation Pi,i+1

are defined as:

Ri+1,i =

[[ri+1,i]3x3 0

0 [ri+1,i]3x3

], Pi,i+1 =

[U [pi,i+1]3x30 U

], (4.3)

where U ∈ �3x3 is the identity operator, ri+1,i is the 3x3 rotation matrix, and pi,i+1 cor-

responds to the skew symmetric matrix corresponding to the vector cross product operator of

position vector pi,i+1. This matrix is described as:

pi,i+1 =

⎡⎣ 0 −pz py

pz 0 −px−py px 0

⎤⎦ (4.4)

Annex 10.1 shows the Matlab-code for the computation of spatial kinematics in �6. The

input of the algorithm corresponds to the wing joint trajectory, whose profile is defined by the

terms: joint positions (q), velocities (qd = q), and accelerations (qdd = q). It returns the spatial

positions (Xb), velocites (Vb) and accelerations (Vb) of the floating body. These spatial terms

are expressed with respect to the body frame {b}.

4.2.3 Wing trajectories and manuevers

Wing joint trajectory profile qi, qi, qi that allow forward and turning flight are extracted from

biological experiments in (20), (22), (23), (24). Figure 4.3a-b describe Cartesian paths for

both turning and forward flight, whereas plots c-d detail wing joint modulation scheme for the

1Detailed information related to rotation matrix properties, DH parameters, homogeneous transformations,

and spatial algebra applied to robotics, can be found in (55) and (91).

54

Page 87: BaTboT: a biologically inspired flapping and morphing bat robot

4.2 Kinematics model

q3

shoulder

q2

q4

q5 q6

q3

q4

q5

q6

elbow

wrist

MCP-III

IVV

(c) (d)

q1

xo

yo zo

yo

flight path

flight path

(a) (b)3 3

banked turn crabbed turn

{o}

{o}

(rolling) (heading)

2

2

q1=q2~=0

xo

yo{o}

xo

yo{o} zo

xo{o}

Figure 4.3: a)-b) Example of Cartesian paths during turning and forward flight, c-d) exam-

ple of wing modulation scheme for wing contraction during upstroke and wing extension during

downstroke. Source: the author.

corresponding maneuvering. It exemplifies how wing joints rotate for contracting and extending

the wing during flight. In the following, wing joint modulation patterns are introduced for both

forward and turning flight maneuvers.

Forward flight

The wingbeat cycle of a bat is composed by two phases: downstroke and upstroke. During

forward flight, both wings should flap symmetrically at the desired wingbeat frequency f . In

(22) a computational model for estimating the mechanics of horizontal flapping flight in bats is

presented. It quantifies (using high speed cameras) how markers placed in the bat wings move

in the three-dimensional space. It also calculates how most relevant joints of the wing rotate

during the wingbeat cycle. Figure 4.4 shows stills of the specimen Cynopterus brachyotis flying

at Brown wind tunnel facility1. Plot-a depicts the beginning of the wing downstroke. The body

1In-vivo experimental data and images of bat flight is copyright of Brown University. Wind-tunnel facility

is at Breuer Lab, School of Engineering http://brown.edu/Research/Breuer-Lab/index.html

55

Page 88: BaTboT: a biologically inspired flapping and morphing bat robot

4.2 Kinematics model

©Brown University

(a) (b)

(c) (d)

q4,5,6

markers

q2 q3

q1

Figure 4.4: Stills of wing kinematics during forward flight: a) beginning of downstroke, b) middle

downstroke, c) end of downstroke/beginning upstroke, d) middle upstroke. Source: (20), (22),

(23), (24).

of the specimen is lined up in a straight line, elbow joint is ∼ 89o. Plot-b depicts the middle

of the wing downstroke, where wings are extended to maximize the area aimed at increasing

lift force, at this state elbow joint is ∼ 60o. Plot-c depicts the end of the wing downstroke in

where the membrane is at maximum camber and the wings remain extended. Elbow joint is

∼ 30o. Plot-d depicts the middle wing upstroke where wings are folded to reduce drag. At this

state elbow joint is ∼ 85o. To know how the angle of the elbow joint rotates (q3) is essential

for the proper modulation of the wing shape. In bats the range of motion of the elbow accounts

for ∼ 90o but most of the rotations measured in (22) have shown an average range of motion

of ∼ 60o.

Kinematics of forward flight is shown in Figure 4.5. It only describes trajectory profiles

of the right wing since the left wing also moves with the same patterns. In (22), recordings

for horizontal flight have been carried out using three cameras placed in the wing-tunnel, as

represented by the cartoon of Figure 4.5 (top). The time change in the X, Y and Z positions

of the wrist joint and third MCP digit to wingtip (with respect to the body frame) were curve-

fitted by polynomials obtained using a least-squares algorithm (KaleidaGraph, version 3.0.5,

Abelbeck Software). In (22) eighth-order polynomials were used to increase precision in the

interpolation of MPC-III joints. Here, the MCP-III digit is considered as a rigid bone connected

to the wrist joint, thus third-order polynomial curves give enough precision for interpolating

56

Page 89: BaTboT: a biologically inspired flapping and morphing bat robot

4.2 Kinematics model

θ[r

ad]

θ[rad/s]

t[s]

θ[rd/s

2]

t[s]

X [m]

Y[m

]

x

y

z

(a) (c)

(b) (d)

downstrokeupstroke

q1 q2 q3 q4 q5 q6

q1 q2 q3 q4 = q5 = q6

q4 = q5 = q6q3q2q1

{1,2}{3} {4,5,6}

wristwingtip

IIIx b[m]

yb[m]

right-wing

cam2

cam1

cam3

q i,Rrad

[]

q i,Rrad

/s[

]q i,R

rad/s2

⎡ ⎣⎤ ⎦

1 2

3

©Brown University

Figure 4.5: (Simulation) Wing kinematics of forward flight (both wings move symmetrically at

wingbeat frequency of f = 2.5Hz): a) Cartesian trajectories of the wrist joint and the wingtip

frame. b) Joint angles. c)-d) Joint velocities and accelerations. For velocities, q3 = q4 (red plot)

and q5 = q6 (purple plot). For accelerations, q3 = q4, q5 = q6. Source: the author.

the Cartesian motion of both wrist and wingtip frames (axis Z does not account since only the

planar motion of both joints is considered). Polynomial values are shown in Table 4.2 whereas

Figure 4.5a shows the Cartesian path of wrist and wingtip frames that are presented by the

polynomials.

To achieve forward flight the wing joints are modulated as described in Figure 4.5b. It shows

the trajectory profile of each wing joint (qi) during a wingbeat cycle (f = 2.5Hz). In the model,

the primary flapping motion (shoulder joint q1) is simply generated as q1 = 60cos(2πf) allowing

the wings to flap an angle range of of 120o at the desired wingbeat frequency f . Similarly the

joint profile q2 = −10cos(2πf−10) allows the wings to rotate forward and backward around the

57

Page 90: BaTboT: a biologically inspired flapping and morphing bat robot

4.2 Kinematics model

Table 4.2: (Forward flight) coefficients of the third-order polynomial curves that describe Carte-

sian paths of wrist and wingtip frames fitted to the xb, yb coordinates of the body frame.

Frame X Y

0.0550 + 0.0475t− 0.0075t3 0.0950− 0.0389t0.0039t3

0.0950 + 0.0250t− 0.0224t2 + 0.0074t3 0.0600− 0.0271t+ 0.0118t2 − 0.0097t3

Wrist 0.1000 + 0.0004t− 0.0240t2 − 0.0118t3 0.0600 + 0.0323t+ 0.0138t2 + 0.0155t3

0.0750− 0.0122t+ 0.0098t2 + 0.0113t3 0.0800 + 0.0132t− 0.0072t2 − 0.0070t3

0.0550− 0.0265t− 0.0033t3 0.0950 + 0.0198t+ 0.0024t3

0.1800 + 0.0406t− 0.0056t3 0.0900 + 0.0124t− 0.0124t3

0.2150 + 0.0237t− 0.0169t2 + 0.0082t3 0.0900− 0.0249t− 0.0373t2 + 0.0122t3

Wingtip 0.1600− 0.0470t+ 0.0129t2 − 0.0121t3 −0.0350− 0.0077t+ 0.0114t2 − 0.0082t3

0.1750 + 0.0150t− 0.0257t2 − 0.0129t3 0.0100 + 0.0396t− 0.0066t2 − 0.0060t3

0.1600 + 0.0022t− 0.0257t2 + 0.0086t3 0.0500 + 0.0444t+ 0.0022t3

gravity axis within an angle range of 20o. On the other hand, the change of wing morphology

is driven by the joint profiles q3...q6. The elbow joint (q3) is key for contracting and extending

the wing during flight. In (22) it was shown that the wingbeat cycle can be partitioned into

approximately 65% of downstroke time required for wing extension and 35% of upstroke time

required for wing contraction, i.e., the wings contract faster to reduce the generation of drag

forces. Here, this approach is achieved by modulating the elbow joint as detailed in Figure 4.5b

(red plot): wings extend in about 0.25s of downstroke time (62.5%) and contract in about 0.15s

of upstroke time (37.5%). Also note how digits III, IV and V (driven by joints q4, q5, q6) open

and close maintaining the minimum angle proportions previously determined in Figure 3.7b:

(downstroke) q4 = 15o, q5 = 65o, and q6 = 128o, (upstroke) q4 = 90o, q5 = 95o, and q6 = 110o.

These proportions contribute to the proper tension of the wing membrane during the process

of extension and contraction.

Turning flight

Kinematics of turning flight is shown in Figure 4.3. In (92), kinematics data of slow turn

maneuvering was recorded as represented by the cartoon of Figure 4.3 (top). To successfully

complete a turn, bats must translate its center of mass along the flight path (i.e. change its

flight direction) and rotate its body around its center of mass to align its body orientation with

the new direction. The magnitude of change in direction of flight is a function of the impulse

(force-time) perpendicular to the original direction of movement. Impulse is the result of the

centripetal force produced by the change of the orientation of the net forces generated by the

58

Page 91: BaTboT: a biologically inspired flapping and morphing bat robot

4.2 Kinematics model

q i,R

(rad)

t(s)

q i,R

(rad/s)

t(s)

¨q i

,R(r

ad/s

2)

t(s)

q i,L

(rad)

t(s)

q i,L

(rad/s)

t(s)

Yb(m)

xb(m

)

¨q i

,L(r

ad/s

2)

t(s)

x

y

zy

z

x

left-wing right-wing

wristwingtip

{0,b}

{1,2}{1,2}{3} {3}

{4,5,6}{4,5,6}

q1 q2 q3 q4 q5 q6

q1 q2 q3 q4 q5 q6

q1 q2 q3 q4 = q5 = q6

q1 q2 q3 q4 = q5 = q6

q4 = q5 = q6q3q2q1

q4 = q5 = q6q3q2q1

(a)

(b) (c) (d)

(e) (f) (g)

yo

xo

zoTop view

Flight path

cam1

cam2

cam3

downstrokeupstroke

©Brown University

markers tracked by cameras allow kinematics quantification

Figure 4.6: (Simulation) Wing kinematics of turning flight at wingbeat frequency of f = 2.5Hz):

a) Cartesian trajectories for the wrist joint and the wingtip frame of each wing (top view). b)-c)-d)

Left wing more contracted: joint angles, velocities and accelerations. e)-f)-g) Right wing more

extended: joint angles, velocities and accelerations. For velocities, q3 = q4 (red plot) and q5 = q6

(purple plot). For accelerations, q3 = q4, q5 = q6. Source: the author.

body and wings. Net forces are the sum of aerodynamic forces (lift and drag) and also inertial

forces (thrust and weight). To turn, bats bank (roll) around its cranio-caudal axis, tilting the

vector of the vertical component of the net aerodynamic force (i.e. lift in level flight) laterally

and towards the center of the turn. The reorientation of net forces produces a laterally oriented

force that drives the bats into a turn.

Based on this foundation, turning flight can be achieved by generating roll momentum by

means of wing contraction and extension, which allows for the displacement the center of mass

of the robot towards the expanded wing. This approach allows the robot to roll (φ). The

59

Page 92: BaTboT: a biologically inspired flapping and morphing bat robot

4.2 Kinematics model

q3

end of wing extension

t =0.4

tR,down = 0.25 tR,up = 0.15

tR,down − tR,up

q4

t =0 t =0.25t =0.15

q3,L

q3,R

q4,L

q4,R

Figure 4.7: Comparison between wing joint trajectories of left and right wings (q3 and q4)

considering a morphing-wing factor of fmc = 0.1. Source: the author.

modulation of the wing joints follows the same trajectory patterns than those presented in

forward flight, however, to achieve the rolling motion one wing should contract and extend is

less proportion compared to the other. This proportion is called the morphing-wing factor fmc.

The influence of the morphing-wing factor can be appreciated in the Cartesian paths of the

wrist and wingtip of the left wing (cf. Figure 4.6a). Cartesian paths differ from the right wing

because the wing joints of the left wing q3,L..q6,L are being modulated with different morphing-

wing factor than the wing joints of the right wing q3,R..q6,R (cf. Figures 4.6b-e). Term fmc is a

number between > 0 and <= 1. If fmc = 1 both wings contract and extend symmetrically in

the same proportion as forward flight does. In Figure 4.6b the left wing has a morphing-wing

factor of fmc = 0.1. Figure 4.7 details the joint motion of q3 and q4 by comparing both left and

right wing profiles from Figures 4.6b-e.

Term fmc puts out of phase and reduces the amplitude of profile q of one wing with respect

to the other. This causes the wings have different contraction and extension periods during

downstroke and upstroke. Note in the case of Figure 4.7, the right wing (q3,R and q4,R) extends

in tR,down = 0.25s and contracts in tR,up = 0.15s but in order to generate inertial forces that

cause de body to roll, the left wing cannot be modulated in the same manner. Thus, the wing

factor is applied to change the modulation of the left wing aimed at increasing the rolling torque

towards the right wing. This is done by q3,L = fmcq3,R (same procedure for q4). This approach

not only reduces the amplitude of q3,L (keeping the left wing more contracted) but also reduces

the extension period of q3,L due to fmc = 0.1 = tR,down − tR,up, making the left wing to

60

Page 93: BaTboT: a biologically inspired flapping and morphing bat robot

4.3 Inertial model

{i}=1

{cm}

si,cm

pi,i+1 {i+1}

body i

body i+1

zi

xiyi

xi+1

yi+1

......i=n=6

......

base

inertial frame

{o}

zo

xoyo

qi

qi+1

{0}

x0y0

z0

zi+1

x

yz

body i

vi{i}

ωifi i

{cm}

......

......

zb

xb

ybbody frame {b}

τφ

τθ

right wing

left wing

i=n=6

{i}=1

{i+1}

zipi,i+1

xi

yi

yi+1

xi+1

zi+1

qi

Fi,Li=n

0

∑ Fi,Ri=n

0

qi+1

τ i

Figure 4.8: Rigid multibody serial chain that composes each wing. Spatial forces of each body

contain both linear fi and angular τi force components stacked into a six-dimensional vector Fi.

These forces are propagated from the wingtip {i} = n to the base frame {0}. Subscripts R,L

denote for right and left wing respectively. The resultant spatial force (FT ) acting on the base

frame {0} is the sum of spatial forces generated by both wings. The inset shows the velocity of a

rigid body i expressed in terms of ωi and vi, and the force acting on a rigid body i expressed in

terms of fi and τi.

decrease aerodynamic forces and therefore to increment the effects of wing inertia modulation

into the production of rolling torque. The rolling torque is mainly caused by the inertial forces

produced by each wing at the center of mass of the body. To analyze the inertial effects of wing

modulation for achieving forward and turning flight, the following section introduces the inertial

model of BaTboT. The inertial model calculates the net inertial forces that are generated at

the center of mass of the robot due to the propagation of forces produced by the wings when

moving using the kinematics profile presented in this section. It also calculates the spatial

accelerations of the body based on applied forces.

4.3 Inertial model

4.3.1 Spatial notation

Equations of Motion (EoM) are formulated based on Newton-Euler formalism using spatial

operator (55). Spatial operators leads to six-dimensional physical quantities that combine the

angular and linear aspects of rigid-body motions and forces. Using spatial algebra notation,

61

Page 94: BaTboT: a biologically inspired flapping and morphing bat robot

4.3 Inertial model

Table 4.3: Spatial operators.

Description 6D operator

Spatial velocities of body i with respect to joint frame {i} Vi = [ωxωyωzυxυyυz ]T ∈ �6x1

Spatial accelerations of body i with respect to joint frame {i} Vi = [ωxωyωz υxυy υz ]T ∈ �6x1

Spatial forces of body i with respect to joint frame {i} Fi = [τxτyτzfxfyfz ]T ∈ �6x1

Spatial rotation from joint frame {i+ 1} to {i} Ri+1,i =

[ri+1,i 0

0 ri+1,i

]∈ �6x6

Spatial translation from joint frame {i} to {i+ 1} Pi,i+1 =

[U pi,i+1

0 U

]∈ �6x6

Spatial translation from joint frame {i} to center of mass {cm} Si,cm =

[U si,cm

0 U

]∈ �6x6

Projection onto the axis of motion of joint i H =[

0 0 1 0 0 0]T ∈ �6x1

Spatial inertia of body i with respect to {cm} Ii,cm =

[Ji,cm 0

0 miU

]∈ �6x6

Inertial tensor of body i with respect to {cm} Ji,cm =

⎡⎢⎣ Ixx 0 0

0 Iyy 0

0 0 Izz

⎤⎥⎦ ∈ �3x3

velocities Vi, accelerations Vi, and forces Fi of a rigid body i are expressed with respect to its

joint frame {i} using six-dimensional vectors, as:

Vi =

[ωiυi

], Vi =

[ωiυi

], Fi =

[τifi

]∈ �6x1 (4.5)

The inset in Figure 4.8 details how the physical components in Eq. (4.5) are expressed in

a rigid body i. In order to propagate the spatial forces produced by the right (Fi,R) and left

(Fi,L) wings onto the base frame {0}, EoM can be recursively solved as:

FT =1∑i=6

Fi,R +1∑i=6

Fi,L ∈ �6x1 (4.6)

FT in Eq. (4.6) is the sum of propagated forces of both wings. The subscripts R and L

refer right and left wing respectively. To propagate Fi,R and Fi,L spatial operators for rotation

Ri+1,i ∈ �6x6 and translation Pi,i+1 ∈ �6x6 in Eq. (4.3) are used. Table 4.3 summarizes all

spatial operators used for the dynamics description of the inertial model. Similarly to Pi,i+1

the operator Si,cm ∈ �6x6 represents a spatial transformation for translation from joint frame

{i} onto the frame located at the center of mass of the body {cm}. The skew symmetric matrix

(si,cm) corresponds to the vector cross product operator of position vector si,cm. Also, each

rigid body with scalar massmi and inertial tensor Ji,cm ∈ �3x3 is composed by a spatial inertial

operator Ii,cm ∈ �6x6 with respect to center of mass (CM) of the body i. The tensor Ji,cm is

composed by moments of inertia (Ixx, Iyy, Izz) which are calculated taking the volume integral

62

Page 95: BaTboT: a biologically inspired flapping and morphing bat robot

4.3 Inertial model

by the square perpendicular distance from the corresponding axis, cf. Eq. (4.7). The term ρ is

the density of the rigid body i. Moments of inertia physically represent the resistance that the

body presents when is accelerating.

Ixx =∫ ∫ ∫ (

y2 + z2)ρdv,

Iyy =∫ ∫ ∫ (

x2 + z2)ρdv,

Izz =∫ ∫ ∫ (

x2 + y2)ρdv,

(4.7)

4.3.2 Equations of motion (EoM)

To solve FT in Eq. (4.6) the term (1∑i=n

Fi) must be derived for each wing, being the subscript

i an indicator of the rigid body: the shoulder, humerus, radius, MCP-III, etc. Spatial forces

are calculated with respect to the joint frame {i} of the rigid body i and then propagated and

oriented towards the next body in the serial chain (cf. Figure 4.8). Spatial operators from

Table 4.3 are used to express EoM, which basically consist on the calculation of forces that

must be applied to the wing-system (Fi) in order to produce a given joint acceleration response

(qi). This is known as the inverse dynamics problem. Therefore, spatial forces and velocities

calculated with respect to the center of mass {cm} of a rigid body i are related with respect to

the joint frame {i} of the body, as:

Fi = STi,cmFi,cm,Vi = STi,cmVi,cm

(4.8)

By definition the spatial force acting on the center of mass of a body i can be expressed by

differentiating the inertial moment (Li,cm = Ii,cmVi,cm) with respect to time, as:

Fi,cm = Li,cm = Ii,cmVi,cm + Ii,cmVi,cm (4.9)

Replacing Eq. (4.9) into Fi in Eq. (4.8):

Fi = STi,cm

[Ii,cmVi,cm + ξ

](4.10)

Where ξ = Ii,cmVi,cm is the gyroscopic force acting on the center of mass frame {cm} of the

rigid body i. The spatial acceleration Vi,cm can be solved by differentiating Vi,cm in Eq. (4.8)

with respect to time, as:

Vi,cm = Si,cmVi,

Vi,cm = Si,cmVi + Si,cmVi(4.11)

63

Page 96: BaTboT: a biologically inspired flapping and morphing bat robot

4.3 Inertial model

Replacing Vi,cm in Eq.(4.11) into Fi in Eq.(4.10) :

Fi = STi,cm

[Ii,cm(Si,cmVi + Si,cmVi) + ξ

]. (4.12)

The term Si,cmVi contains Coriolis effects. The spatial forces in Eq. (4.12) can be recur-

sively propagated along the wing by considering the spatial operators for rotation Ri+1,i and

translation Pi,i+1. Therefore, spatial forces in Eq. (4.12) are backward propagated, from the

wingtip (i = n = 6), to the base frame {0} (i = 1), as:

1∑i=6

Fi = Ri+1,iPTi,i+1Fi+1 + STi,cm

[Ii,cm(Si,cmVi + Si,cmVi) + ξ

]= Ri+1,iP

Ti,i+1Fi+1 + STi,cmIi,cmSi,cmVi + STi,cmIi,cmSi,cmVi + STi,cmIi,cmVi,cm

(4.13)

The expression Ri+1,iPTi,i+1Fi+1 allows for the projection of spatial forces along the serial

chain of bodies that compose the wing structure. Also note from Eq. (4.13) that the term

STi,cmIi,cmSi,cm refers to the spatial inertia Ii calculated with respect to the joint frame {i} by

applying the parallel axis theorem. Equation (4.13) is rewritten as:

1∑i=6

Fi = Ri+1,iPTi,i+1Fi+1 + IiVi +

[IiSi,cm + Ii

]Vi. (4.14)

To complete the solution of the inertial model in Eq. (4.14), the set of spatial velocities (Vi)

and accelerations (Vi) are also recursively calculated as:

Vi = Pi,i+1RTi+1,iVi−1 +Hqi

Vi = Pi,i+1RTi+1,iVi−1 + Pi,i+1R

Ti+1,iVi−1 +Hqi + Hqi

(4.15)

The term H allows for the projection of the joint trajectory profile qi,qi onto the axis of

motion. Wing joint trajectories are presented in Figure 4.6. Algorithm 1 describes how to

solve FT in Eq. (4.6) and Annex 10.2 details the Matlab-code for solving Algorithm 1. The

inputs of the algorithm are the wing joint profiles (qi, qi, qi) and the kinematics velocity (Vb) and

acceleration (Vb) of the body calculated from kinematics algorithm in Annex 10.1. Algorithm

1 returns the spatial forces FT and accelerations of the floating body Vb that are produced by

FT . Both terms are with respect to the body frame {b} due to the applied rotation R0,b. It

rotates Fi that is with respect to the frame {0} by R0,b = rz0(−π/2)rx0(π) (cf. Figure 4.2).

Vb =

(Ib +

[R0,b

0∑i=6

(Ri+1,iPTi,i+1Ii,R)+R0,b

0∑i=6

(Ri+1,iPTi,i+1Ii,L)

])−1

FT (4.16)

64

Page 97: BaTboT: a biologically inspired flapping and morphing bat robot

4.3 Inertial model

Algorithm 1 Inertial model computation

initialize:

k = 1, V0 ←[

φ θ ψ 0 0 0], V0 ←

[φ θ ψ 0 0 9.81

]while k ≤ 2, do –for each wing–:

——————————————————-

1. Forward recurrence: spatial velocities and accelerations

for i = 1 → n do

Vi ← Pi,i+1RTi+1,iVi−1 +Hqi

Vi ← Pi,i+1RTi+1,iVi−1 + Pi,i+1R

Ti+1,iVi−1 +Hqi + Hqi

end for

——————————————————-

if (aerodynamics are included) do Fi+1 ←[

0 0 0 −FD 0 FL

]Telse do Fi+1 ←

[0 0 0 0 0 0

]T2. Backward Recurrence: spatial forces

for i = n → 1 do

Fi ← Ri+1,iPTi,i+1Fi+1 + IiVi +

[IiSi,cm + Ii

]Vi //forces

τi ← HTFi //torques

if i! = n do Ii + [Ri+1,iPTi,i+1]Ii+1[Ri+1,iPi,i+1] //inertia propagation

end for

if k == 1 do Fi,R ← Fi, Ii,R ← Ii //(R): right wing

else do Fi,L ← Fi, Ii,L ← Ii //(L):left wing

k ++

end while-do

——————————————————-

3. Compute body forces with respect to body frame {b}FT ← R0,bFi,R +R0,bFi,L

4. Compute body inertia Ib with respect to body frame {b}Ib +

[R0,b

0∑i=6

(Ri+1,iPTi,i+1Ii,R)+R0,b

0∑i=6

(Ri+1,iPTi,i+1Ii,L)

]5. Compute body acceleration caused by FT with respect to body frame {b}Vb =

(Ib +

[R0,b

0∑i=6

(Ri+1,iPTi,i+1Ii,R)+R0,b

0∑i=6

(Ri+1,iPTi,i+1Ii,L)

])−1

FT

Return FT , Vb

4.3.3 Rolling and pitching torques

To achieve forward and turning flight, section 4.2.3 showed how to kinematically take advan-

tage of the morphing-wing modulation aimed at changing the wing shape. Dynamically, these

changes in wing shape induce effective forces (torques) at the center of mass of the robot that

allow the body to rotate. Here, the inertial model in Eq. (4.14) is used to define rolling (τφ)

and pitching torques (τθ) components that induce forward and turning flight respectively. Both

terms are extracted from the components of FT as:

τφ =[1 0 0 0 0 0

]FT

τθ =[0 1 0 0 0 0

]FT

(4.17)

Rolling and pitching torques from Eq. (4.17) are with respect the body frame {b} as shown

65

Page 98: BaTboT: a biologically inspired flapping and morphing bat robot

4.4 Aerodynamics model

FT(x)

FT(z)

FL

FD

Fnet

Fnet

Figure 4.9: Free-body diagram of a bat in accelerating flight, indicating the aerodynamic and

gravitational forces that accelerate the center of mass (COM). Lift is perpendicular to the direction

of flight whereas drag and thrust are parallel to the direction of flight. The net force produced

can be decomposed into net force components parallel and perpendicular to the direction of flight

(see inset). The parallel component corresponds to the net thrust. Thus, measurements of the

acceleration of the COM would directly reflect the net forces acting on it. Source: (24).

in Figure 4.8. Simulations in section 4.6.2 show estimations about the influence of wing inertia

into the production of rolling and pitching torques.

4.4 Aerodynamics model

Figure 4.9 shows free-body diagram of a bat in accelerating flight, indicating the aerodynamic

and inertial forces that the center of mass of the robot. Inertial forces i.e., thrust and weight

components, can be directly modeled using the inertial model (FT ) from Eq. (4.14) by ex-

tracting the force components of the Zb and Xb axes respectively. Aerodynamics forces i.e., lift

and drag components, can be directly quantified by measuring the force produced by the real

robotic platform. It is more accurate to have an identified model of aerodynamics forces rather

than a simulated model using fluid dynamics theory. Experiments to quantified aerodynamic

forces will be carried out in Section 7.3 of Chapter 7. For instance, this section provides the

basic foundations to measure aerodynamics and net forces.

4.4.1 Lift and drag forces

Lift is perpendicular to the direction of flight whereas drag and thrust are parallel to the

direction of flight. Figure 4.10a shows the testbed, which shows BaTboT (in the inset)1 mounted

1Due to aerodynamics modeling depends strictly on real experiments aimed at quantifying lift and drag

coefficients, this section gives a brief insight into the experimental tests to carry out using the final BaTboT

66

Page 99: BaTboT: a biologically inspired flapping and morphing bat robot

4.4 Aerodynamics model

(b)

CL

CD

(a)

AoA

drag (FD)thrust (FT,x)

FT,z

lift (FL)

{o}

zo

yoxoFnet

Force sensor

AoA [deg]

Figure 4.10: (Experimental) aerodynamics identification –wind-tunnel–: a) the bat-robot is

mounted on top of a 6-DoF force sensor from which both lift FL and drag FD forces are experi-

mentally calculated as a function of the airflow speed and angle of attack (AoA), b) Lift and drag

coefficients (CL, CD) calculated from measured lift and drag forces. Source: the author.

in the wind tunnel, on the end of a supporting sting that defines the angle of attack (AoA).

The robot is mounted on top of a 6-DoF force sensor1 from which both lift CL and drag CD

coefficients are calculated from the force measurements of lift and drag forces.Typical results

are shown in Figure 4.10b. Both lift (L) and drag (D) force components are measured directly

from the force sensor by applying Eq. (4.18).

L = FLcos(AoA)− FDsin(AoA)D = FDcos(AoA) + FLsin(AoA)

(4.18)

Then, the lift and drag coefficients are computed as:

CL = 2L(ρV 2airAb)

−1 CD = 2D(ρV 2airAb)

−1 (4.19)

Where the term ρ = 1.20Kgm3 is the air density, Vair is the airspeed, and Ab = 0.05m2 is

the planform area of the wing. Using Eq.(4.18) and (4.19) identified aerodynamic forces can be

incorporated into the inertial model of the robot, as detailed in Algorithm 1 (cf. before step 2).

At the beginning of the force propagation (i = 6), the term Fi+1 = 0 in the absence of external

forces, i.e., accounting only inertial contribution. In case aerodynamics forces are included into

the inertial model, the term Fi+1 would correspond to:[0 0 0 −FD 0 FL

].

prototype. Further experiments with detailed analysis of aerodynamics behavior is presented in Chapter 7

section 7.3.1force sensor Nano17 transducer ATI Industrial Automation, 0.318 gram-force of resolution, http://www.

ati-ia.com/

67

Page 100: BaTboT: a biologically inspired flapping and morphing bat robot

4.4 Aerodynamics model

4.4.2 Net forces

As explained by (24), to accelerate during forward flight, any flying organism must produce a net

aerodynamic force to counteract gravity and overcome drag. This force can be decomposed into

a net force component in the direction of flight that corresponds to the difference between thrust

and drag, i.e. net thrust, and a perpendicular component that corresponds to the difference

between lift and weight (cf. Figure 4.9). During forward, steady flight, the average lift over the

course of a wingbeat must equal body weight, and average thrust must equal drag. However,

unlike airplanes, flying organisms cannot continuously generate constant lift and thrust because

of the oscillatory nature of flapping, so instantaneous force generation varies across the wingbeat

cycle. As a consequence, a flying bat will accelerate and decelerate throughout a wingbeat, even

during steady-state flights where average acceleration is zero over the complete cycle. In (24) is

demonstrated that in slow flight bats generate a net forward force during the upstroke: During

upstroke, the upward and backward acceleration of the wings will produce an inertial force

that will move the body forward and downward with respect to the downstroke. This force

will produce a forward-oriented component, or inertial thrust (FT,x) during upstroke. During

downstroke, the downward and forward acceleration of the wings will produce an inertial force

that will move the body backward and upward while keeping the position of the center of mass

constant. The horizontal component of this inertial force will produce negative inertial thrust

during downstroke (−FT,x). So bats produce positive inertial thrust during the upstroke motion

of the wings mainly due to the body moves in opposition to the flapping direction in order to

conserve momentum. Bats properly modulate wing kinematics to take a maximum advantage

of wing inertia on the production of body accelerations. These accelerations are produced by

the net forces (Fnet), which are calculated with respect to the body frame {b}, as:

Fnet =([

0 0 0 1 0 0]FT − FD

)+(FL − [

0 0 0 0 0 1]FT

)(4.20)

Inertial forces are likely to be significant in bats because the mass of the wings comprises

a significant portion of total body mass, ranging from 11% to 33% and because wings undergo

large accelerations. In this specie, wings’ mass accounts for 31% of the total mass, while in the

robot wings’ mass accounts for 37% of the total mass. In this thesis it will be demonstrated

the significant impact of wing modulation kinematics into the production of net forces. By

properly controlling wing modulation, both inertial (FT ) and aerodynamics forces (FL, FD) will

be efficiently produce aimed at generating large net forces (Fnet) during forward flight.

68

Page 101: BaTboT: a biologically inspired flapping and morphing bat robot

4.5 SMA muscle-like wing actuation

SMA

spring

SMA1

SMA2

rj rj

lsma

SMA1

SMA2-+

I1 Fsma-1

Fsma-2

I2τ 3

τ 3τ 3

q3

SMA1

-+

I1 Fsma-1τ 3

lsp −k ∗ lsp(a) (b)

Biceps

Triceps

Figure 4.11: SMA actuation configurations: a) (top) SMA joint with bias spring concept,

(medium) mechanic implementation, (below) SMA-spring model representation. b) (top) SMA

antagonistic joint, (medium) mechanic implementation that mimic how bicep and tricep muscles

operate, (below) antagonistic pair of SMAs model representation. In both configurations SMA

wires extend along the humerus bone of BaTboT wings, acting as artificial muscles that pull the

elbow joint q3. SMA pulling forces (Fsma) produce a joint torque (τ3) that rotates the elbow,

pulling in the ”fingers” to slim the wing profile on the upstroke. Source: the author.

4.5 SMA muscle-like wing actuation

Recently, Shape Memory Alloys (SMAs) have opened new alternatives and the potential of

building lighter and smaller smart actuation systems (43), (44), (45), (93), (46), (94). To closely

mimic the morphing-wing apparatus of bats, extremely light wires with negligible volume acting

as artificial muscle fibers seem to be an adequate solution. To this purpose SMAs wires able to

contract upon electrical heating have been used as artificial biceps and triceps that allow the

wing to contract and extend by means of elbow and wrist joints rotation. This section provides a

comprehensive review of SMA modeling, highlighting the thermo-mechanical properties behind

NiTi material aimed at defining equations that partially govern phenomenological behavior, i.e,

shape memory effect, hysteresis, strain, stress, etc. To this purpose a phenomenological model

is presented in order to assess the performance and limits of SMA technology for the application

at hand.

69

Page 102: BaTboT: a biologically inspired flapping and morphing bat robot

4.5 SMA muscle-like wing actuation

4.5.1 SMA actuation configurations

Upon heating SMA wires use the one-way shape memory effect to generate force and motion.

Because SMA actuators can only contract in one direction, it is necessary to provide a biasing

force to return to the neutral position. This can be accomplished using a bias spring (cf. 4.11a),

or another SMA element in an antagonistic arrangement (cf. 4.11b). In practice the former is

mostly used, but the latter gives further control range.

In the SMA actuators with bias spring arrangement, only one SMA is heated and cooled, so

the hysteresis effect has quite a significant influence on control performance. This configuration

has the advantage of having no slack issues over the SMA or stress when recovering to the neutral

position. On the contrary the antagonistic configuration, which heats one actuator while the

other cools, can reduce the hysteresis effect, as experimentally demonstrated by (73), and (82).

The advantage of this configuration can be appreciated from a control perspective since instead

of providing passive biasing force or motion, both directions can be actively controlled. This

increases the range of controllable actuation at expense of adding a phenomenon called the two-

way shape memory effect which is mainly produced when the wires extend upon cooling. The

passive SMA wire can develop a few millimeters of slack as it cools, which consequently affects

the accuracy of the control. The two-way shape memory effect becomes even more problematic

in the antagonistic arrangement of SMA actuators, and may lead to slower response and even

wire entanglement. Since this is a control problem, section 6.5 in Chapter 6 will approach these

issues.

For instance, the following subsection focuses on describing how to properly model thermo-

mechanical properties of SMAs and its use as muscle-like actuators. This allows to assess and

explore the limits to increase SMA performance in terms of actuation speed and output torque

by keeping the input power below the limits of overheating.

4.5.2 SMA phenomenological model

SMAs exhibit an unique thermomechanical property due to the phase transformation of the

material, from austenite phase to martensite phase and vice versa. These transformations

mainly occur due to changes in temperature and stress. Extensive research has been devoted

to model these properties. Tanaka in (71) was one of the pioneers to study a stress-induced

martensite phase transformation, proposing an unified one-dimensional phenomenological model

that make use of three state variables to describe that process: temperature T , strain ε, and

70

Page 103: BaTboT: a biologically inspired flapping and morphing bat robot

4.5 SMA muscle-like wing actuation

martensite fraction ξ. His main contribution was to demonstrate that the rate of stress is a

function of strain, temperature and martensite fraction rates. Later, Brinson (78) improved

on Tanaka’s model by separating the calculation of the martensite fraction into two parts,

one induced by stress and the other one induced by temperature. This issue allowed for the

description of the shape memory effect at low temperatures.

Elahinia (56),(70) proposed an enhanced phenomenological model compared to the previous

ones, and also addressed the nonlinear control problem. This model was able to better describe

the behavior of SMAs in cases where the temperature and stress states changed simultaneously.

Their model was verified against experimental data regarding a SMA-actuated robotic arm. As

a result, the phenomenological model was able to predict SMA behavior also under complex

thermomechanical loadings. Further experiments were also carried out in (95).

In this thesis, Elahinia’s phenomenological model (70) has been used for assessing the limits

of SMA operation. The model consists of four parts: i) heat transfer, ii) mechanics model,

iii) forward/reverse phase transformation, and iv) kinematics model. As shown in Figure 4.11,

the input of the model is the electrical current Isma to drive the SMA and the output is the

elbow joint rotation q3 that is produced by the strain rate of the SMA when pulling the joint.

This model allows for determining proper parameters to safe overload SMA performance without

compromising physical damage to the shape memory effect or overheating issues when subjected

to high amount of input power. Simulations in section 4.6.4 are carried out for characterizing

overload SMA actuation response that would be fundamental for control tuning and experiments

with the real platform.

Heat transfer model

The SMA wire heat transfer equation consists of electrical (Joule) heating and natural convec-

tion:

msmacpT = I2smaRsma − hcAc (T − To) (4.21)

SMA NiTi wires have a diameter of 150μm, a mass per unit length of msma = ρπr2j where ρ

is the density of wire, 2rj is diameter of wire, Ac = π2rj is circumferential area of the unit length

of the wire, cp is specific heat, Isma is applied electrical current, Rsma is electrical resistance

per unit length of the wire, T is temperature of the wire, To is the ambient temperature, and

hc is the heat convection coefficient. Although in Eq. (4.21) is assumed that hc and Rsma are

both constant, a detailed experimental analysis on the resistance variation of the SMAs can be

71

Page 104: BaTboT: a biologically inspired flapping and morphing bat robot

4.5 SMA muscle-like wing actuation

found in Section 6.4. Using Eq. 4.21 is possible to model how the NiTi wire would heat upon

electrical current and by removing the term I2smaRsma (heating power), the equation can be

also used to model how the NiTi wire cools in the absence of heating power.

Mechanical model

SMA mechanical model was firstly introduced by Tanaka in (71). It relates stress rate (σ) with

temperature rate (T ) as:

σ =θs−Ω(Af−As)

−1

1−Ω(Af−As)−1Cm

T (4.22)

Where θs corresponds to the thermal expansion factor of the wire, Ω is the phase transfor-

mation factor, Af , As are the austenite final and initial temperatures and Cm is the effect of

stress coefficient on martensite temperature. Also, the strain rate (ε) during heating phase can

be calculated as:

ε = σ−θsT−ΩξEA

(4.23)

Where EA is the austenite the Young’s modulus and ξ is the phase transformation rate

which is presented in the following.

Phase transformation model

The reverse transformation equation that describes the phase transformation from martensite

to austenite during heating is:

ξ = ξm2 [cos (aA (T −As) + bAσ) + 1] (4.24)

where ξ is martensite fraction that has a value between 1 (martensite phase) and 0 (austenite

phase). The terms aA = π(Af − As)−1 and bA = −aAC−1

A are the curve-fitting parameters

of the phase transformation. Also, the forward transformation equation describing the phase

transformation from austenite to martensite during cooling is:

ξ = 1−ξa2

[cos (aM (T −Mf ) + bMσ) +

1+ξa2

](4.25)

Where aM = π(Mf −Ms)−1 and bM = −aMC−1

M are the curve-fitting parameters, where

Mf ,Ms are the martensite phase final and initial temperature respectively.

72

Page 105: BaTboT: a biologically inspired flapping and morphing bat robot

4.5 SMA muscle-like wing actuation

Table 4.4: Parameters for SMA phenomenological model

Variable Model Parameters Value [unit]

Temperature

Heating: msma,Rsma, Isma 1.14 × 10−4 [Kg], 8.5 [Ω]

msmacpT = I2smaR − hcAc (T − To) Ac 1.76 × 10−8[m2

]

Cooling: hc 150[Jm−2◦C−1s−1

]

(T ) msmacpT = −hcAc (T − To) Cp 0.2[KcalKg−1◦C−1

]

Stress (σ)

Heating: Ω −1.12 [GPa]

σ =θs−Ω

(Af−As

)−1

1−Ω(Af−As

)−1Cm

T θs 0.55[MP◦aC−1

]

Cooling: Cm,Ca 10.3[MP◦aC−1

]

σ =θs−Ω

(Ms−Mf

)−1

1−Ω(Ms−Mf

)−1Ca

T As,Af ,Ms,Mf 68, 78, 52, 42[◦C]

Strain (ε)

Heating:

ε =σ−θsT−Ωξ

EAEA 75 [GPa]

Cooling: EM 28 [GPa]

ε =σ−θsT−Ωξ

EM

FM (ξ)

Heating: ξm, ξa 1, 0 [dimensionless]

ξ =ξm2

[cos

(aA (T − As) + bAσ

)+ 1

]aA 0.31

[◦C−1]

Cooling: aM 0.31[◦C−1

]

ξ =1−ξa

2

[cos

(aM

(T −Mf

)+ bMσ

)+

1+ξa2

]bA, bM −0.03

[◦C−1]

Kinematics model

The kinematic model describes the relationship between SMA strain rate ε with the angular

rate of the elbow joint q3:

q3 = lsmaε(2rj)−1, (4.26)

The rotation of the elbow joint (q3) can be calculated by integrating Eq. (4.26) with respect

to time, being lsma the length of the SMA wires, and rj the radius of the joint (cf. Figure

4.11b).

SMA phenomenological algorithm

Table 4.4 summarizes the parameters used for the simulation of the thermo-mechanical equa-

tions. Further details on the values assigned to most coefficients can be also found in (56) and

(70).

Algorithm 2 describes the procedure for computing the thermo-mechanical equations of the

SMA phenomenological model. Also Annex 10.4 shows the Matlab-code for Algorithm 2. Inputs

are the electrical currents (Isma) to drive each SMA actuator. The output is the elbow joint

rotation q3. Step 1 computes the temperature rate of the SMA (T ) based on current inputs.

It allows for the evaluation of overheating when the SMAs are subject to high values of input

currents. In step 2 SMA stress rate is calculated (σ). Whether the temperature on SMA wires

increases dramatically, stress would also increase allowing for the identification of over-stressing

73

Page 106: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

Algorithm 2 SMA thermo-mechanical computation

1. Compute SMA temperature rate: T ← m−1smac

−1p (I2smaRsma − hcAc (T − To))

2. Calculate the SMA stress rate upon heating: σ ← θs−Ω(Af−As)−1

1−Ω(Af−As)−1

CmT

3. Calculate phase transformation rate (differentiating Eq. (4.24) with respect to time):

ξ ← − ξm2

[sin (aA (T −As) + bAσ) + (aAT + bAσ)

]4. Compute SMA strain rate upon heating: ε ← σ−θsT−Ωξ

EA

5. Compute kinematic model of joint motion rate: q3 ← lsmaε(2rj)−1

6. Integrate q3

7. Return q3

limits. In step 3 the calculation of the phase transformation allows for the analysis of the

hysteresis loop when SMAs are transitioning from martensite to austenite and vice versa. In

step 4 SMA strain is calculated allowing for the estimation of elbow joint rotation rate.

4.6 Simulation and experimental results

Simulations and experiments are aimed at analyzing three issues:

1. Wing torques for actuation: It quantifies the required torques to properly move the joints

of the wing skeleton. It allows for the characterization of actuators (simulation).

2. Body torques for maneuvering : It quantifies the influence of wing inertia into the pro-

duction of rolling and pitching torques for maneuvering. It allows for the assessing of

forward and turning flight. Also, it is useful for attitude control tuning and requirements

for proper wing modulation (simulation and experimental).

3. SMA actuation limits : It explores the limits to safe overload the response of the SMA

actuators by defining the maximum value of input electrical current that achieves the

fastest rotation speed of the elbow joint (simulation).

An open-loop Matlab-based simulator has been implemented using the SimMechanics tool-

box of Simulink1. Figure 4.12 details the main modules that compose the simulator. One key

advantage of the simulator consists on the possibility of using the CAD model of the robot

exported directly from SolidWorks2. This allows to include the mechanical assembly together

kinematics and dynamics properties of the robot into the simulation environment. The sim-

ulator will be extended to a closed-loop architecture in Chapter 6.

1http://www.mathworks.es/products/simmechanics/index.html2http://www.solidworks.com/

74

Page 107: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

4.6.1 Open-loop simulator

This simulator is composed by the following modules (cf. Figure 4.12):

• Wing trajectory generator : It generates trajectory patterns (qi, qi, qi) for each joint of the

wings at specified wingbeat frequency f . These patterns are similar to those shown in

Figures 4.5 and 4.5.

• Kinematics module: it places frames of references into the mechanical assembly of the

robot. It also maps the wing trajectory patterns to the corresponding wing joint. Spatial

operators for rotation and translation are calculated (cf. Annex 10.1).

• Dynamics : It contains the inertial model (cf. Algorithm 1). It calculates the required

torques (τi) to produce the desired wing motion patterns (qi, qi, qi). Also, it calculates

the spatial inertial forces (FT ) that are generated at the center of mass of the robot due

to the motion of both wings. This allows for quantifying the influence of wing inertia for

maneuvering.

• SMA actuation: It maps position commands for elbow and wrist rotation (q3) to out-

put torque (τ3). The output torque is generated by calculating the required amount of

input heating power (I2smaRsma) to drive SMA strain to desired elbow rotation (q3), cf.

Algorithm 2.

• Mechanics module: It contains the mechanical assembly and properties of robot’s CAD

exported directly from SolidWorks (cf. section 5.1 in Chapter 5 to further details on CAD

design).

4.6.2 Wing torques for actuation

To characterize the actuators for wing actuation, the inertial model in Algorithm 1 is used. It

allows for the estimation of wing torques (τi) based on desired wing trajectory profiles (qi).

Figure 4.13 describes how wing torques are applied to the bat robot. Effective forces τ1, τ2 are

the torques applied to the shoulder joint. The former allows for the primary flapping motion

(q1) whereas the latter for the forward/backward rotation of the wings about the gravity axis

(q2). The effective force τ3 is the torque applied to the elbow joint. It allows for the contraction

and extension of the wing via elbow rotation (q3). Also, τ4, τ5 and τ6 are the torques applied

to the wrist joint which rotate the digits to maintain wing membrane tension. Further details

on actuator(s) selection based on the simulation data presented in this section can be found in

75

Page 108: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

Wing

traje

ctory

gen

erat

or m

odule

SMA

actu

ation

mod

ule

Kine

mat

ics, m

echa

nics,

dyna

mics

mod

ules

User

ope

ratio

n

3D e

scen

ario

Figure

4.12:Sim

Mechanicsopen

-loopsimulatorfordynamicsandSMA

actuation.Sourc

e:th

eauth

or.

76

Page 109: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

q3

q4,5,6

q2

q7

xb

yb θ

φ

xb

ybτφ

τθ

z2

z1x1

y1

y2x2

z3

x3

y3

x4,5,6

y4,5,6

z4,5,6τ1

τ 2

τ 3

τ 4,5,6

Inertialmodelqi τ i

aerodynamics

robotparameters

Figure 4.13: Wing torques denoted as τi correspond to the effective forces that the each joint i

requires to rotate as defined by trajectory profiles qi (see upper inset). Torques are with respect

to the joint frames {i} assigned by modified DH convention. The lower inset shows the model for

the estimation of torques as a function of the joint profiles. The inertial model in Algorithm 1

takes into account robot parameters/constrains and aerodynamic loads. Source: the author.

Table 4.5: (forward flight) wing torques as a function of the wingbeat frequency f .

f = 1.3Hz f = 2.5Hz f = 5Hz f = 10Hz

|τ1|[Nm] 0.00016 0.0006 0.0024 0.0096

|τ2|[Nm] 0.00037 0.0013 0.0054 0.022

|τ3|[Nm] 0.0007 0.0027 0.011 0.0438

|τ4,5,6|[Nm] 0.00027 0.0010 0.004 0.0162

Chapter 5 section 5.3.1.2. For instance, the following simulation are aimed at characterizing

wing torque requirements for: i) forward flight maneuver, ii) turning flight maneuver, and iii)

energy-demanding kinematics pose.

Required torques during forward flight

Here, wing profiles qi are applied according to the kinematics trajectories defined to forward

flight in Figures 4.5. The response of the inertial model is shown in Figure 4.14. Because during

forward flight both wings flap symmetrically describing the same wing profile, the torque values

shown in Figure 4.14a apply for both left and right wings. This wing torque profile corresponds

to a wingbeat frequency of f = 5Hz, which is the typical flapping frequency during steady

flight. From the figure is easy to appreciate how the elbow joint (red plot) demands for about

77

Page 110: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

0.01

0.02

0.03

0.04

0

-0.01

-0.021.3 2.5 5 10

−τ 1−−τ 2−−τ 3−−τ 4,5,6−

(a)

(b)

Figure 4.14: (Simulation) a) required wing torques (τi) during forward flight. Both wings flap

symmetrically describing the wing profiles qi shown in Figure 4.5b (wingbeat frequency f = 5Hz),

b) Increments of wingbeat frequency cause the wing torques to increase. The plot shows maximum

peak values of wing torques Source: the author.

4.5-times more applied torque than the flapping motion powered by the shoulder joint (from

2.4mNm to 11mNm). This is caused by gyroscopic forces generated during the flapping motion

which directly affect the in-plane contraction and extension of the wings. These results provide

a first insight about how much output torque the SMA actuators should produce to drive the

elbow and wrist rotations (τ3 + τ4,5,6). Also, to assess and quantified how wing torques scale

as a function of the wingbeat frequency (f), Figure 4.14b details the response of the inertial

model considering different wingbeat frequencies ranging from 1.3Hz up to 10Hz. In plot-b

the x-axis corresponds to four sets of frequencies. Each set contains the wing torques of each

joint. As f increases, note how the wing torque profile also increases due to the increment

of angular accelerations of each joint. Plot-b is useful because provides an insight about how

the wing torque profile scale up when f increases, being not only useful for the selection of

actuators but also for control tuning. An scaling factor denoted as Winertial−factor has been

quantified from the results in plot-b. Several simulation results have been carried out to confirm

that the wing torque profile scales by a factor of ∼ 4 when the wingbeat frequency is doubled

78

Page 111: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

1.3 2.5 5 10 1.3 2.5 5 10

−τ 1−−τ 2−−τ 3−−τ 4,5,6−

−τ 1−−τ 2−−τ 3−−τ 4,5,6−

Wingbeat frequency f [Hz] Wingbeat frequency f [Hz]

(a) (b)

Figure 4.15: (Simulation) a) required wing torques (τi) during turning flight: a) left wing,

b) right wing. The upper plots show the difference between the wing torque profile that each

wing requires to describe the trajectory profiles qi shown in Figure 4.6b-e respectively (wingbeat

frequency f = 5Hz). The lower plots show how increments of wingbeat frequency cause the wing

torques of each wing to increase. Source: the author.

(Winertial−factor = 4). Table 4.5 reports the numerical data.

Required torques during turning flight

Here, wing profiles qi are applied according to the kinematics trajectories defined for turning

flight in Figure 4.6. The response of the inertial model is shown in Figure 4.15. Contrary to

forward flight where both wings are modulated symmetrically, turning flight maneuver requires

the robot to modulate each wing differently. The modulation profile for each wing was previously

presented in Figures 4.6b-e. To goal behind this approach is to generate body torques at the

center of mass of the robot that result on producing body angular accelerations that allow for

turning left or right. In this simulation the left wing has been modulated with a morphing wing

factor of 0.1 which corresponds to the kinematics profile shown in Figures 4.6a. This causes the

left wing to be more contracted with respect to the right one. This causes the center of mass to

displace towards the right wing and therefore generating rolling torque around the body frame

{b} (cf. 4.16b, lower plot).

This approach works a shown in Figure 4.16a. Up to t = 2.5s both wings are modulated

symmetrically with the same wing trajectory profile qi, which causes net inertial forces to cancel

79

Page 112: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

..

..

..

yb

xbφ

(a) (b)

Figure 4.16: (Simulation) effects of wing modulation on the production of body torques and

angular accelerations at wingbeat frequency of f = 1.3Hz a) wing modulation profile qi for the

left wing (upper) and right wing (lower), b) (upper) body torques (τθ, τφ, τψ) produced at center

of mass, (medium) body angular accelerations (θ, φ, ψ) produced at center of mass, and (lower)

attitude response of the robot (θ, φ, ψ). Values are with respect to the body frame {b}, Source:the author.

each other at the center of mass of the body, cf. Figure 4.16b (upper plot). As a consequence zero

body accelerations in roll (φ) and yaw (ψ) are produced, cf. Figure 4.16b (medium plot). Now

from t > 2.5s, the right wing reduces the amplitude of the joint modulation by a morphing-wing

factor of fmc = 0.5 (2.5 < t ≤ 3.8) and then with fmc = 0.1 (t > 3.8), cf. Figure 4.16a (lower

plot). As a consequence rolling and yawing torques are being produced, causing a dominant

body angular acceleration in roll that makes the body to rotate towards the expanded left wing.

It is important to mention that body torque components in Figure 4.16b (upper plot) have been

extracted from the spatial body forces FT calculated by the inertial model using Algorithm 1.

Likewise body angular accelerations in Figure 4.16b (medium plot) have been extracted from

the spatial body accelerations Ab, whereas attitude response in Figure 4.16b (lower plot) is

calculated from the double integration of angular accelerations with respect to time. Further

simulations in Figure 4.17 show how body torques (upper plots), body acceleration (medium

plots) and attitude response (lower plots) are increased when the wingbeat frequency increases

up to f = 10Hz.

Also, Table 4.6 details numerical values for the characterization of torque requirements for

each joint of the wing qi. It specifies how much torque should be applied depending on the

80

Page 113: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

..

..

..

..

..

..

..

..

..

0.2

0.1

0

-0.1

0.15 0.2

0-0.20

-0.150

(a) (b) (c)

Figure 4.17: (Simulation) effects of wing modulation on the production of body torques and

angular accelerations at wingbeat frequency of a) f = 2.5Hz, b) f = 5Hz, and c) f = 10Hz.

(upper) body torques,(medium) body angular accelerations, (lower) attitude response. Source:

the author.

Table 4.6: Characterization of actuators: wing torque requirements for flapping at 10Hz and

morphing1 at 2.5Hz

Parameter q1 q2 q3 q4,5,6

Joint rotation range2 qi[deg] 120 25 60 45, 30, 18

Joint speed q[deg/ms] 120/50 25/50 60/200 45, 30, 18/200

Torque3 τi[Nm] 0.0096 0.022 0.0027 0.001

Torque (±5% tolerance) τi[Nm] 0.010 0.023 0.0028 0.00105

1 Morphing motion is slower due to SMA limitations, cf section4.6.4.2 Joint rotation ranges applied during forward flight, cf. Figure 4.5b.3 It contains aerodynamics loads experimentally quantified in Figure 4.10.

specific joint speed and its kinematics range. This characterization will be useful for the selection

of the flapping motor and also for the selection of SMA morphing-wing actuators. The following

section details how to properly generate rolling and pitching torques for maneuvering based on

changing wing morphology. It allows for the experimental quantification of the influence of

wing inertia on the production of body forces and accelerations. Also, simulation results are

contrasted against experimental measurements of body forces aimed at validating the inertial

model in Algorithm 1.

81

Page 114: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

4.6.3 Body torques for maneuvering: experiments for inertial model

validation

Back in section 4.4.2 it was explained that net forces Fnet (cf. Eq. (4.20)) varies during the

wingbeat cycle in part due to the inertial thrust (forward-oriented component: FT,x, cf. Figure

4.9) changes from upstroke to downstroke and vice versa. During downstroke, the downward

and forward acceleration of the wings produce an inertial force that moves the body backward

and upward, which produce negative inertial thrust. On the contrary, the upstroke motion

produces positive inertial thrust mainly due to the body moves in opposition to the flapping

direction in order to conserve momentum. This issue has been simulated and experimentally

proven in Figure 4.18.

Figure 4.18a details the end of the upstroke motion whereas Figure 4.18b the ends of the

downstroke motion. Upper plots show experiments carried out to measure 6D forces that are

generated by the wings during the wingbeat cycle. The same testbed used in aerodynamics

experiments from Figure 4.10 was used for inertial measurements1. Lower plots show simulation

results using the SimMechanics simulator shown in Figure 4.12. As mentioned, the simulator

contains the inertial model from Algorithm 1 that allows for the computation of inertial forces

FT and net forces Fnet. Firstly it can be noted from Figure 4.18 (lower plots) how the primary

flapping motion of the wing generates forward and backward pitch motion (θ) measured about

the body frame {b} and with respect to the inertial frame {o}. Pitch motion is generated by

the pitching torque (τθ) produced when the wings are placed forward and backward the body.

Quantification of pitching torque (τθ) and also rolling torque (τφ) are consigned in Figure 4.19.

To generate forward pitching torque (τθ) both wings must be positioned towards the body

(see inset from Fig. 4.19a). The blue plot corresponds to the experimental values of τθ measured

by the force sensor, whereas the black plot corresponds to the response of the inertial model

in Algorithm 1. The simulated pitching torque has been extracted from the spatial force FT

as defined in Eq. (4.17). The disparity in results between the simulation model and the

experimental results are due to two factors. Firstly, mechanical asymmetries due to small

fabrication errors have caused the robot to generate a small component of pitching torque

(τθ =∼ 0.06Nm) even when q2 = 0o (from t = 0s to t = 2s in Fig. 4.19a). The solution

of this problem requires the incorporation of an attitude controller that regulates the pitch

motion. Secondly, the disparity in the amplitude of the oscillations is due to non-modeled

1force sensor Nano17 transducer ATI Industrial Automation, 0.318 gram-force of resolution, http://www.

ati-ia.com/

82

Page 115: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

UpstrokeDowntroke

zb

xb

yb

{b} zb

xb FT,x

zb

xb

FT,x

(a)

(b)

force sensor

CMCM

Upstroke Downtroke

(a)

Figure 4.18: Effects of the flapping motion of the wings on the position of the center of mass (CM)

and accelerations of the body: a) end of upstroke motion, b) end of downstroke motion. Upper

plots depicts the testbed for experimental measurements of six-dimensional inertial forces (FT )

and lower plots depicts simulations for the computation of inertial forces (FT ). Values for both

measured and simulated FT are consigned in following Figure 4.19. To conserve momentum, the

body moves in opposition to the flapping direction. During upstroke, the upward and backward

acceleration caused by the flapping motions of the wings produce an inertial force (red circled

arrow) that moves the body forward and downward with respect to the downstroke. This force

produces a forward-oriented component (FT,x), or inertial thrust (green solid arrow). Contrary,

during the downstroke negative inertial thrust is produced. Source: the author.

physical quantities, such as mechanical frictions. This issue may cause the simulation values to

be larger in amplitude in comparison to the experimental readings.

To generate rolling torque (τφ) that allows the robot for turning left or right one wing must

be contracted while the other is extended (see inset from Fig. 4.19b). In this experiment similar

disparity problems to those observed from the pitch test scenario are presented. However, note

from both Fig. 4.19a-b that from t = 2s to t = 4s the model and the experimental values tend

to stabilize about the same bias value of torque. In plot-a the bias value of τθ =∼ 0.31Nm,

whereas in plot-b the bias value of τφ =∼ 0.11Nm.

Figure 4.19c-d shows how bias values of τθ and τφ scale up when the wingbeat frequency

83

Page 116: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

[Nm]

2 [deg]

[Nm]

3 [deg]

qyb

(a) (b)

(c) (d)

1.3Hz 2.5Hz5Hz

1.3Hz 2.5Hz5Hz

xb

q

xb

yb

q

xb

yb

q3 (deg)q deg

t (s) t (s)

τ θ(Nm)

τ θ(Nm)

τ φ(Nm)

τ φ(Nm)

q deg q deg, q deg

q deg, q deg

q

xb

yb

Figure 4.19: (Experimental) wing inertia contribution on forward and turning flight: a) sim-

ulation model VS experimental results of pitching torques τθ (f=5Hz), b) simulation model VS

experimental results of rolling torques τφ (f=5Hz), c)-d) quantification of wing inertia contribution

into the generation of τθ and τφ at different wingbeat frequencies f . Source: the author.

increases. Table 4.7 details the numerical results of the scaling factors for both pitching and

rolling torques. It has been observed that the influence of wing and body mass (mt = 0.125Kg)

on the production of inertial torques is proportional to m1/ft , ∀f > 0. Biological studies in

(24) show experimental evidence about the implication of this relationship in terms of inertial

and aerodynamics contribution. For the application at hand, mt has accounted for ∼ 50% of

pitching torque production, and ∼ 20% for rolling torque production. By adding the membrane

to the wings, the rest is contribution of aerodynamic forces.

Other interesting tests in Figures 4.20 and 4.21 show how the production of pitching and

rolling torques respectively is affected when considering non-biologically inspired wing joint

references but instead step-input signals that drive wing joint rotations.

The set M1 corresponds to the inertial torques generated by using the biological trajectory

84

Page 117: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

Table 4.7: Wing and body mass influence in the generation of τθ and τφ (mt = 0.125kg)

Scaling factor f = 1.3Hz f = 2.5Hz f = 5Hz

Pitch1(0.5m1/ft )

Experimental 0.11Nm 0.2Nm 0.31Nm

Simulation 0.1Nm 0.21Nm 0.33Nm

Roll2(0.2m1/ft )

Experimental 0.04Nm 0.07Nm 0.11Nm

Simulation 0.04Nm 0.08Nm 0.13Nm

1 Values from Fig. 4.19c2 Values from Fig. 4.19d

y

x

Figure 4.20: (Simulation) pitching torque response (τθ) for different positions of q2 (wings rotated

forward/backward the body) and for different wingbeat frequencies f . The set M1 contains bio-

inspired wing modulation shown in Figure 4.5b whereas set M2 corresponds to step-input signals

that rotate each joint of the wing due to the range defined in Table 4.6 (cf. joint rotation range).

Source: the author.

patterns, whereas the set M2 using step-commands patters. In the first place one can note a

significant differences between the torques values obtained from both simulations. Table 4.8

details the numerical results. The kinematics modulation of the wing joints has a key impact into

the production of rolling and pitching torques for body rotation. It is precisely the geometry of

the wings that causes different inertial and aerodynamic responses depending of the modulation

profile that changes its shape during flapping. In most of the cases, note how the percentage

of torque increase is greater than 50%, which conclusively stands for the importance of using

bio-inspired modulation profiles presented in Figures 4.5 and 4.6. By including control methods

that ensure wings are always modulated with the proper bio-inspired patterns, section 7.2.1 in

Chapter 7 will introduce further experimental results about the importance and influence of wing

inertia modulation for more efficient flight.

85

Page 118: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

y

x

Figure 4.21: (Simulation) rolling torque response (τφ) for different positions of q2 and wingbeat

frequencies f . The set M1 contains bio-inspired wing modulation whereas set M2 corresponds to

step-input signals that rotate each joint of the wing. Source: the author.

Table 4.8: Advantage of using bio-inspired wing joint trajectories on the production of pitching

and rolling torques).

Trajectory set f = 1.3Hz f = 2.5Hz f = 5Hz 10Hz

τθ[Nm] (Pitching1)

M2 0.08 0.15 0.2 0.3

M1 0.11 0.2 0.35 0.55

% of increase 37% 34% 75% 84%

τφ[Nm] (Rolling2)

M2 0.01 0.025 0.06 0.09

M1 0.02 0.04 0.09 0.15

% of increase 100% 60% 50% 67%

1 Values from Figures 4.20 for q2 = 20o

2 Values from Figures 4.21 for q2 = −20o

4.6.4 SMA-muscle limitations

This section presents simulation results that allow for the assessment of SMA limitations in

terms of: i) actuation speed, and ii) power consumption. Both issues are fundamental for

tackling the SMA control problem aimed at driving morphing-wing motions. Using the SMA

phenomenological model previously introduced in Algorithm 2, simulations are focused on ex-

amining how an SMA wire respond upon high values of input current (Isma). This allows for

the definition of input power limits to overload SMA nominal operation. Therefore nominal and

overloaded SMA operation are evaluated. Nominal operation mode is defined by the manufac-

86

Page 119: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

0 0.1 0.2 0.3 0.4 0.50

7

14

21

28

35

t[s]

[deg

]

0 0.1 0.2 0.3 0.4 0.5

50

100

150

200

t[s]

T[de

g]

0 20 40 600

50

100

200

0.02 0.03 0.04 0.05 0.06 0.0750

100

150

200

[MPa

]

@350mA@215mA@175mA@1A

@215mA@175mA@350mA@1A

Lower Nominal Overloaded Overheated

1

1

2

2 3

34

4

3

(d)

T [°

C]

cooling

heating350mA

ISMA =550mA

0

15

30

45

60

75

ISMA =350mAISMA =175mAISMA =1A

q3 [d

eg]

t [s]

(a)

[MPa

]

I=350mAI=175mAI=550mAI=1A

0.04 0.03 0.02 0.0150

200

heating

350mAI =350mA(b)

50454035302520

20

40

60

80

100T [C]

% A

uste

nite

20 25 30 35 40 45 5020

40

60

80

100

% M

artensite

Cooling

HeatingMs

Af

AS

Mf

(c)

Nominal(350mA)

(a)

T [°C]

t [s]

25 30 35 40 4520 5020

40

60

80

100

150

100

20

40

60

80

100

ISMA =350mAISMA =175mAISMA =550mAISMA =1A

Figure 4.22: (Simulation) SMA phenomenological model response at different current profiles:

a) Joint rotation based on SMA strain. b) Temperatures on the SMA wire, c) Hysteresis-loop for

the nominal operation mode (Isma = 350mA), d) SMA strain VS stress.

turer of the NiTi SMA wire (Migamotor, http://store.migamotors.com/, nanomuscle-nm706

super linear muscle wire actuator), cf. section 5.3.4 in Chapter 5 to further details on SMA

actuators supplied by Migamotors. This muscle-like actuator consists of several short strips of

SMA wire attached to opposite ends of six metal strips stacked in parallel. Each SMA segment

pulls the next strip about 0.67mm relative to the previous strip, and the relative movements

sum to make a stroke of 4mm. The SMA mechanical configuration shown in Figure 4.11b allows

an elbow rotation range of q3 = 60o (the radius of the elbow joint is rj = 0.004m).

In summary, under nominal operation the SMA wires have a strain of 4% and are able to

contract in 300ms while pulling 12.23gram − cm (torque at joint of 0.0012Nm) by applying

an input power of 0.26W (Isma = 175mA) at 3V . In this thesis alloys of 150μm of diameter,

with length of Lsma = 0.035(×6)m and nominal electrical resistance of Rsma = 8.5Ω are used.

SMA parameters are described in Table 4.4. On the other hand, overloaded mode refers to

increases in input electrical currents that drive faster joint rotation rates. Because using larger

input currents may cause physical damage of the shape memory effect, simulation analysis is

convenient.

Figure 4.22 shows the response of the phenomenological model (cf. Algorithm 2) under

different values of input heating currents Isma, ranging from 175mA up to 1A. Figure 4.22a

87

Page 120: BaTboT: a biologically inspired flapping and morphing bat robot

4.6 Simulation and experimental results

shows the angular rotation profiles (e.g. elbow’s joint q3) obtained for each input current.

Figure 4.22b shows increases on SMA wire temperature (T ) corresponding to the results shown

in plot-a (cooling temperatures are detailed in the inset). Figure 4.22c details the hysteresis loop

of the SMA wire at input current of Isma = 350mA (states from Austenite to Martensite and

viceversa). Finally, Figure 4.22d depicts the strain (ε) versus stress (σ) curves corresponding

to the results shown in plot-b.

To assess how fast the elbow joint can rotate the kinematics model from Eq. (4.26) is used.

It relates the strain rate of the SMA wire (ε) with the rotation speed of the elbow joint (q3).

As a first attempt, an input current of 175mA is applied, being this value, the nominal input

current suggested by the manufacturer of the SMA actuators (96). Note that under this value,

the joint rotates ∼ 60o in 400ms, resulting too slow for the application at hand. As expected,

by increasing the input current up to 1A, faster angular motions were achieved (up to ∼ 60o in

75ms), however, the SMA wire finish austenite temperature (see Figure 4.22b) was dramatically

increased up to Af = 150oC, being this value about twice higher than the upper value defined

for the simulation Af = 78oC (safe temperature limit of the alloy, see Table 4.4). This issue

clearly illustrates an overheating problem. Overheating causes an increase in the cooling time

of the SMAs, which makes the actuators slower. It also can cause physical damage to the shape

memory effect.

In order to classify the SMA’s operation modes upon the responses of input heating currents

Isma, four modes of SMA operation have been defined: 1-lower, 2-Nominal, 3-Overloaded, 4-

Overheated/overstressed. Both nominal and overloaded modes are feasible targets to pursuit

with the experiment, e.g., by applying an input current around 550mA, the joint is able to

rotate 60o in 100ms while keeping the maximum limits of temperature and stress below the

limits of overheating, i.e., above the limits of the finish austenite temperature Af .

As long as the input heating current Isma is smaller than a maximum allowed current Imax,

the SMAs will be safe, and the angular speed resulting from this input current could be set as

a feasible target speed to pursuit in the experiment. For instance, simulation results suggest

Imax =∼ 550mA. Table 4.9 summarizes numerical values to safe overload SMA operation.

Nominal mode allows for a morphing-wing frequency of 1.3Hz whereas overloaded mode allows

for a morphing-wing frequency of 2.5Hz. These values are defined by the simulation results of

elbow joint angular speed carried out in Figure 4.22a.

88

Page 121: BaTboT: a biologically inspired flapping and morphing bat robot

4.7 Remarks

Table 4.9: Limits for overloading SMA operation

Parameter Theoretical1 Simulation2

Nominal (f = 1.3Hz)

Max. input current Isma [mA] 175 350

Max. input heating power uheating [W ] 0.26 1.04

Resultant joint speed q3 [deg/ms] 60/300 60/200

Resultant output torque τ3 [Nm] 0.0012 0.0007

Overloaded (f = 2.5Hz)

Max. input current Isma [mA] N/A 550

Max. input heating power uheating [W ] N/A 2.57

Resultant joint speed q3 N/A 60/100

Resultant output torque τ3 [Nm] N/A 0.0027

1 Nominal values provided by Migamotor’s model NanoMuscle RS-70-CE 1131 (96). Values do not consider external

load. Overloaded values are not provided.2 Simulation results from Figures 4.22a (resultant joint speed at input current), and Table 4.5 (resultant output

torques). Values do consider the wing-skeleton load.

4.7 Remarks

This chapter has presented the formulation of kinematics, dynamics, aerodynamics and SMA

actuation via modeling frameworks that are a key step towards the design/fabrication, control,

and flight experiments using BaTboT.

The most important contributions of this chapter can be summarized as follows:

• Flight maneuvers were kinematically formulated in Figures 4.5 and 4.6. It has shown how

the bio-inspired kinematics of wing joint modulation has allowed the definition of forward

and turning flight. As a result, wing joint trajectory profiles were determined.

• An inertial model formulated in Algorithm 1 has allowed the assessment and quantifica-

tion about how wing inertia affects robot’s maneuverability, and the key role of proper

wing modulation aimed at the production of rolling and pitching torques for turning

and forward flight. It has also allowed the characterization of torques requirements for

actuator selection.

• A phenomenological model formulated in Algorithm 2 has allowed the assessment and

quantification of limits to overload SMA operation by preventing the wires to overheat

and overstress.

The following chapter introduces the design and fabrication process of BaTboT and its

components.

89

Page 122: BaTboT: a biologically inspired flapping and morphing bat robot

5

BaTboT design and Fabrication

The rich diversity of mechanisms of animal flight can provide abundant inspiration for

engineered design.

5.1 The general method for BaTboT’s design

Flapping flight is the single most evolutionarily successful mode of animal locomotion: there are

today over 1200 species of bats, more than 10000 living species of flying birds, and somewhere

between millions and tens of millions of species of flying insects. Understanding how animals fly

is not only central to providing insight into the biological world; the rich diversity of mechanisms

of animal flight can provide abundant inspiration for engineered design.

This chapter introduces the design and fabrication process of BaTboT, the first bat-like

micro aerial vehicle with highly articulated wings. This chapter is not intended to cover

performance of the mechanical designs presented, but only the design process,

main functions and the criteria involved. Refer to Chapters 4 (modelling) and 6

(Control) for detailed experiments regarding the mechanical approaches introduced

herein. In this chapter, one of the most important challenges to be tackled concerned to

the design criteria to specify morphological and actuation parameters. Because the novelty

of the robot and the lack of information regarding design issues, this thesis was meant to

design BaTboT from the analysis of a specific bat specimen aimed at carefully mimicking each

detail related to morphology, biomechanics, kinematics and even the muscle-like wing actuation

system. Chapter 3 already presented this biological analysis by concluding the chapter with

key issues or foundations to the design process. Here, those foundations are bring back and

incorporated into the design framework that brings BaTboT to live.

90

Page 123: BaTboT: a biologically inspired flapping and morphing bat robot

5.1 The general method for BaTboT’s design

wing markers

1. Biological study: kinematics extraction of

wings' motions

2. Biological inspiration:Morphology, physiology

and design criteria

3. Fixed-wing design:Flapping testbed and membrane

testing. Flapping actuation selectionElbowjoint

Carpus/wristjoint

MCP-III digitshoulder

4. Articulated-wing design:Metal tendons provide wrist

motion

air-bubbles within the membrane detected!!

SMA-based muscles

5. Morphing-wing mechanism:Antagonistic design of SMA-based muscles as actuators.

Plagiopatagium skin(0.1mm silicon wing

membrane)

R/C transmission link49MHz

Antenna

motor+electronics

SMAmorphing muscles

6. Proper membrane design and testing:inertial force measurements of joint

torques.7. Robot assembly + electronics:

control encoded.

9. Flight tests

Re-designing process and adjustments

8. Wind-tunnel measurements.Wind-tunnel testing of inertial and

aerodynamics forces. Control response testing.

Figure 5.1: Design process of BaTboT. Source: The author.

91

Page 124: BaTboT: a biologically inspired flapping and morphing bat robot

5.2 Prototype characteristics

5.2 Prototype characteristics

This section briefly gives an insight about the design process previously mentioned and the

main components of the robot. The following sections describe seven major steps that are key

into the design process of the robot.

5.2.1 Design process

The design process is depicted in Figure 5.1. Steps 1 and 2 highlight the key biological founda-

tions concluded from Chapter 3, section 3.3.2. It details how biological based parameters are

used for the CAD design of the robot. Step 3 shows the first approaches for the design and

fabrication of the wing skeleton and the membrane. It details the first problems and solutions

to transform a rigid skeleton into a articulated wing. Step 4 and 5 present the final design of

the articulated wing that incorporate the muscle-like actuation system using Shape Memory

Alloys. In step 6, the final design and fabrication process of the wing membrane is presented

and attached to the wing skeleton. Experiments are conducted for testing the elasticity and

cambering properties of the silicone-based membrane. Finally, step 7 shows the components of

the robot, from electronics and sensors to the final assembly of BaTboT.

5.2.2 Components and weight distribution

Figure 5.2 details the most relevant components of the robot and their weight distribution.

Components are classified into two categories: i) structural components such as the body,

shoulder, wings and ii) electronic components such as the actuators, arduino-board, battery

and sensors. Structural components have been designed using SolidWorks CAD and fabricated

via ABS (Acrylonitrile butadiene styrene) 3D printing process. All components or parts of the

robot are hollow plastic structures that minimise the overall weight. Structural components

are shown in Table 5.1.

Wing components (i.e, shoulder, humerus, radius, digits and wing membrane) have been

carefully designed based on morphological parameters analysed from the specimen Pteropus

poliocephalu. In general, BaTboT is about 50% smaller than the true specimen’s size, but it

maintains the biological proportions of body and wing size. This is because the wing-to-body

mass ratio play an important role for maneuvering aspects that have to be included in the

flight control. Because in nature bats modulate solely wing inertia to produce pitch and roll

moments, wing mass is crucial. As shown in Table 5.2 the specimen has a wing-to-body mass

92

Page 125: BaTboT: a biologically inspired flapping and morphing bat robot

5.2 Prototype characteristics

Monday, July 9, 12

humerusbone

leg

shoulder

MPC-III

radiusbone

IV

V

body

arduinonano

IMU

LiPobatery

SMA driver

SMA muscles

spring bias

motor

elbowjoint

wristjoint

(a)

(b)

[grams]

Figure 5.2: a) weight distribution of main components, b) detailed view of main components.

Source: The author.

ratio of 0.404, whereas the robot has a wing-to-body mass ratio of 0.410. This closer correlation

enables high levels of performance in terms of pitching and rolling torque production.

93

Page 126: BaTboT: a biologically inspired flapping and morphing bat robot

5.2 Prototype characteristics

Table 5.1: Main structural components of BaTboT. CAD design

CAD Description

Body. It is the main structure of the robot with a mass of

38g. Body width is 7.8cm, a required length to achieve the

angle range of motion of the wings from the rotation axis of

the flapping DC-motor. In the middle of the body is located

the flapping motor-case (in grey colour), which consists of a

reinforced structure that holds the motor. The reinforcement

is because the larger torsional forces produced during flapping.

The size of the structure has been set to allow the motor an

specific x, y, z position to drive the flapping of the wings. In

addition the structure contains an extra space for the second

axis of rotation. Section 5.3 details on the requirements of the

flapping-wing mechanism. All the compartments inside the

body work as hosts for the electronic components, as shown

in Figure 5.1b.

.

Shoulder. It is a structure with a mass of 3g and carefully

designed to fit in the body. The robot has two shoulders

to drive the flapping motion of each wing. Both wings are

attached to each shoulder which provides the corresponding

angle of attack to each of them (AoA = 9o). Section 5.3

details the flapping-wing mechanism.

.

Humerus bone. It is a structure with a mass of 4g and a length

of 5.5cm. This artificial bone connects the shoulder joint with

the elbow joint and hosts the SMA muscle-like actuators that

act as biceps and triceps. In the specimen this bone has a

length of 11cm (cf. Table 5.2). The artificial humerus is al-

most half the size of the biological one. The idea behind this

choice concerns to reducing the size of the robot’s wingspan

to the half aimed at fabricating a small-scale vehicle.

.

Radius bone. It is a structure with a mass of 2g and a length

of 7cm. This artificial bone connects the elbow joint with the

wrist joint. Inside the bone, metal tendons are extended to

allow for the motion of the wrist joint. Subsection 5.3.4 details

the morphing-wing mechanism. Also, this bone is about half

the size of the the biological one (cf. Table 5.2).

.

Digits MCP-III,IV,V. Three artificial digits with a respec-

tive mass of 0.3, 0.2, 0, 2g complement BaTboT’s wings. The

MCP-V acts as fifth digit and thumb. The thumb section has

a length of 1.76cm and it is cambered 5o. This contributes in

cambering the attached membrane. Also, the MCP-III is com-

posed by two sections, one acting as second digit; the shortest

with 5.42cm and the third digit; the longest with 10.54cm.

Digits are all connected to form the wrist joint that rotates

(open and close) as a function of the elbow joint. This sub-

actuation is driven by the metal tendons.

.

94

Page 127: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

Table 5.2: Comparison of morphological parameters between the specimen and BaTboT

Parameter [unit] Bat 1 Robot

Total mass mt [g] 98 125 (with battery)

Extended wings’ length B [m] 0.462 0.245

Body width lm [m] 0.07 0.04

Body mass mb [g] 41 38

Wing mass2 mw [g] 16.58 15.6

Wing-to-body mass ratio mw/mb 0.404 0.410

Body inertia tensor diagonal Ib [gcm2] – [1, 0.07, 0]

Extended wings’ span S = lm + 2B [m] 0.99 0.53

Extended wing’s area Ab [m2] 0.069 0.05

Humerus length lh [m] 0.11 0.055

Humerus average diameter 2rh [m] 0.0055 0.006

Humerus inertia tensor diag. Jh,cm [gcm2] – [0.03, 0.37, 0.93]

Humerus position vector to CM sh,cm [m] – [0.0275, 0, 0]

Radius length lr [m] 0.145 0.070

Radius average diameter 2rr [m] 0.0042 0.005

Radius inertia tensor diag. Jr,cm [gcm2] – [0.07, 0.92, 0.37]

Radius position vector to CM sr,cm [m] – [0.035, 0, 0]

Plagiopatagium membrane thickness [m] 0.0002 0.0001

1 Morphological parameters of the specimen extracted from (22).2 wing mass is composed by: humerus (4x2g), radius (2x2g), MCP-III,IV,V (0.7x2g), SMA (1.1x2g).

5.3 BaTboT mechanics

This section details the seven steps mentioned in the Design Process flow depicted in Figure 5.1.

Step 1: Biological study was already carried out in Chapter 3: From bats to BaTboT: Mimicking

biology. Data extracted from in-vivo experiments presented from the biological literature was

analysed and used as a framework for the i) selection of the specie to mimic and ii) BaTboT

design. Regarding the first point, several species were evaluated in terms of flight performance

and morphological characteristics. The evaluation focused on the Rousettus aegyptiacus speci-

men, with similar physiological and aerodynamical behaviour than the Pteropus poliocephalu.

Kinematics, dynamics and aerodynamics data was then compared. Regarding the second point,

Chapter 3 was concluded with key bio-inspired geometrical parameters for modeling and design

(cf. Table 3.5) and a biological-based framework for design aimed at achieving the best flight

performance found in nature (cf. Table 3.6). In the following, both criteria are consolidated for

the final design parameters of BaTboT.

95

Page 128: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

humerus bone

radius bonethumb

leg

elbow joint

wrist joint

SMA musclesbiceps, triceps

shoulder joint

wing membrane

MCP-III

MCP-IV

MCP-V

(a)

(b)

Figure 5.3: a) Each morphological aspect of BaTboT design has been carefully approached to

its biological counterpart, b) Cartoon of typical wingtip trajectory described during a wingbeat

cycle. Source: The author.

5.3.1 Step 2: Design criteria

This subsection reviews key criteria, from BaTboT design to final fabrication. For design,

bio-inspired parameters found in the specimen are used, whereas for fabrication, the inertial

model of the robot is used. The inertial model (inverse/forward dynamics model) is useful for

the estimation of actuators and general performance. The following subsection details on both

issues.

5.3.1.1 Bio-inspired parameters

In Figure 5.3a, it can be really appreciated how BaTboT design has been carefully conceived

aimed at mimicking as close as possible the morphology of the biological counterpart. Each arti-

ficial bone and skeleton component (cf. Table 5.1) has been shaped and dimensioned according

to the morphological data studied and extracted from the specimen.

As mentioned before, the robot is in general half the size of the bat, however,

skeletal proportions in terms of lengths and weights have been attempted to

maintain. This is extremely important since most of the dynamics and

aerodynamics performance relies on these proportions, specially in the ratio

between wing size/mass and body mass.

96

Page 129: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

Morphological parameters are completely described in Table 5.2, whereas efficient param-

eters for design as a function of body and wing mass (mt) are presented in Table 3.6. In the

following, these parameters are highlighted, explaining how they affect dynamics and aerody-

namics behaviour. Dynamics is quantified in terms of rolling (τφ) and pitching (τθ) torque

production, which are key variables that enable efficient manoeuvrability. Aerodynamics is

quantified in terms of lift and drag production.

• Wingspan: it defines the area of the wings, larger wingspans induce larger lift forces but

also produces larger drag forces. Because BaTboT is able to modulate its wings (change

the area), minimum and maximum wingspan is defined as 0.41 − 0.46m, generating a

maximum wing area of 0.064m2. BaTboT wings have to be massive aimed at accounting

sufficient inertial forces to maneuver. However heavy wings might constraint the robot to

properly flight. The tradeoff between wing size and weight is defined as the ratio between

wing mass and wing-length. It has been observed from the specimen this ratio is about

0.35g/cm. Because BaTboT has the half of the wingspan, wing-mass to wing-length ratio

is about twice the biological one: 0.65g/cm. This difference is partly compensated with

the total mass of the robot.

• Wing-to-body mass: this ratio is crucial for the proper impact of wing inertia into

the production of body accelerations. In the specimen this ratio is 0.404 and BaTboT

has a wing-to-body mass of 0.410. This implies a correlation of 98.51%. This criterion

is perhaps the most important and influential for robot design, since the hy-

pothesis of this thesis is related to the quantification of wing inertia and its

key influence in robot’s manoeuvrability and net force production. To achieve

this closer correlation in the wing-to-body mass ratio between the specimen and BaTboT,

both body and wing skeleton of the robot was carefully shaped in the CAD (SolidWorks)

to finally match the desired mass (ABS plastic properties). Refer to Table 4.7 and Figures

4.19, 4.20 to observe how wing inertia influence was quantified in both simulation and

experimental testing.

• Kinematics of flight: it is mainly represented by flapping frequency (f). From Table 3.6

it has been observed that biological bats do not decrease the flapping frequency lower than

1.47Hz (wingbeat period of 0.68s). In addition, wings must have a wingstroke amplitude,

or flapping motion amplitude of 157.8o and a minimum angle of attack of ∼ 7o. Both

issues drastically influence the production of aerodynamic forces.

97

Page 130: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

• Aerodynamics of flight: flapping motion amplitude and wing’s angle of attack are not

the only kinematic variables that influence on aerodynamics behaviour. Wing camber

is also important. No-cambering constraints the proper flow of air through the wing,

whereas excessive cambering generates turbulences that increase drag and decrease the

lifting surface. Maximum wing camber is set to 0.15. Also, aerodynamics is affected by

the way wings are modulated, in other words, how to properly control the morphing-wings

to achieve the higher lift-to-drag ratio.

As shown in Table 3.6, the previous parameters vary depending on the total mass (mt). For

the application at hand, mt = 0.125Kg (including battery). In case of fabricating a bigger or

smaller robot, the design framework introduced in Table 3.6 determines how to design the most

important components of the robot that enable the kinematics, dynamics and aerodynamics

behaviour found in nature. Chapter 7 will assess and demonstrate via direct experimentation

whether these design parameters really induce the high performance expected.

5.3.1.2 Actuators

BaTboT features a hybrid drive, partially actuated by a DC motor which drives the primary

flapping motion and SMA actuators which drive the morphing-wing motion. Both actuators

have been selected based on the simulation results given by the computation of the inertial

model. Table 4.6 contains the numerical data. From the table, the joints of interest are: q1

(flapping) and q3 (morphing). In the practise, the other joints are under-actuated, which means

they have no direct actuation. Table 5.3 summarises the torque requirement criteria for the

proper selection of both flapping and morphing-wing actuators.

The flapping-wing actuator is a miniature gearmotor with dimensions (0.94” x 0.39” x

0.47”), high-power brushed DC motor with 50 : 1 gear ratio and 3mm-diameter D-shaped

output shaft. It has a mass of 9g, a rotation speed of 625RPM at 6V and 100mA free-run. It

supports a maximum current of 1.6A stall with an output torque of 1.1kg − cm. This brushed

DC gearmotor is intended for use at 6V , though in general, this kind of motor can run at

voltages above and below the nominal voltage, so it should comfortably operate in the 3− 9V

range (rotation can start at voltages as low as 0.5V ). Lower voltages might not be practical,

and higher voltages could start negatively affecting the life of the motor. Later in subsection

5.3.2.1, the response of the motor is characterised for the application at hand (flapping-wing

mechanism).

98

Page 131: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

Table 5.3: Actuators

Actuator Criteria3 Datasheet

Flapping1 Torque: 0.010Nm,

speed: 120/50(deg/ms)

(f = 10Hz)

Torque: 1.1kg − cm (0.1078Nm),

speed: 625RPM (f = 10.41Hz)

Morphing2 Torque: 0.0028Nm,

speed: 60/200(deg/ms)

(f = 2.5Hz)

Torque4 : (0.0012Nm), speed4 :

60/300(deg/ms) (f = 1.6Hz)

1 http://www.pololu.com/catalog/product/998.2 http://store.migamotors.com.3 Torque requirements have ±5% tolerance and contain aerodynamic loads quantified in Figure 4.10.4 Torque and speed values of SMA migamotor actuators correspond to their nominal response. They have to be

overloaded aimed at matching the requirements.

(a)

(b)

Figure 5.4: a) Pin-out and Circuit Diagrams for SMA migamotor actuator model NM706-Super,

b) actuator dimensions. Source: The author.

The morphing-wing actuator consists of several short strips of Shape Memory Alloy (SMA)

wire attached to opposite ends of six metal strips stacked in parallel. Each SMA segment pulls

the next strip about 0.67mm relative to the previous strip, and the relative movements sum to

make a stroke of 4mm. Figure 5.4 details the circuit connection diagram and the dimensions of

the actuator in both extended and contracted states. Table 5.4 details the pinout connections

of the SMA actuator.

99

Page 132: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

Table 5.4: Circuit pinout data for SMA migamotor actuator.

Terminal Description I/O

EXTENDED signal indicating status of full SMA extension O

CONTRACTED signal indicating status of full SMA contraction O

CTRL current signal controlling contraction I

VDC Applied voltage I

GND ground –

The maximum applied voltage VDC is 20V . When CTRL> 2.5V , the actuator contracts and

the CONTRACTED line is pulled to VDC . If an external return force, such as an antagonistic

SMA or spring produced an external pull force, the EXTENDED signal is pulled to VDC and

the unit is prepare to contract once again. For the application at hand, pull-down resistors

are necessary because the EXTENDED and CONTRACTED signals are monitored by the

arduino-board. The actuator supports driven electrical currents up to 2A, however, applying

larger currents during continuos operation may cause overheating problems, fatigue issues or

even physical damage of the shape memory effect. Driven currents have always to be monitored.

The actuator optimal operates in ambient temperatures ranging from −70oC to +75oC and

it has a cycle life of 106+. Datasheet can be downloaded at http://www.migamotors.com/

Downloads.html.

One of the key features of SMA actuators rely on their capability to act like sensors. The

inner SMA wires have an electrical resistance of 8.5Ω that decreases when temperature on the

wire increases. Several properties of the SMAs change as it undergoes the martensite phase

transformation. Among these properties is the resistivity that decreases as the temperature of

the wire increases and hence its phase transforms to austenite. The resistance changes almost

linearly with the strain, which in turn causes a liner relationship with the angle joint rotation

q3. This resistance-angle relationship is linear because the martensite fraction is kinematically

coupled to the rotation, and the martensite fraction is what drives the resistance changes.

Further information of both SMA actuation and sensing capabilities are described in subsection

5.3.4.

5.3.2 Step 3: Fixed-wing design

The design process of BaTboT (cf. Figure 5.1) beings with the design and fabrication of a

fixed-wing that closely match the skeleton structure of the specimen. A testbed was developed

100

Page 133: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

1mmsilicone-membrane

half-bodyflapping-testbed

Angular acceleration VS Power

q1 deg/ s2⎡⎣ ⎤⎦

450 5 10 15 20 25 30 35 40

Pow

er [W

]

2

0

0.4

0.8

1.2

1.6

2.4

Figure 5.5: (left) half fixed-wing flapping testbed, (right) experimental quantification of power

requirements for flapping at maximum f = 10Hz. Source: The author.

aimed at testing the real requirements for flapping motion and to discover how to attach an

artificial membrane to the wing skeleton. Figure 5.5 shows the flapping testbed.

5.3.2.1 Flapping-wing mechanism

To power flapping motion, biological bats have four downstroke muscles: pectoralis major, sub-

scapularis, part of serratus anterior, and part of the deltoid. Together, these muscles constitute

about 12% of the bat’s body weight. The upstroke is powered by the deltoid, trapezius, the

rhomboids, infraspinatus, and supraspinatus (cf. Figure 2.5). These powerful muscles give

an insight about the huge power requirements for flapping flight in bats. In fact, Figure 2.3d

showed that the average power-to-mass ratio (wing mass) is about 0.04W/g, for a specimen

with a mass of 100g flapping at 10Hz. This gives a power requirement of 4W . Comparing the

biological data against the measurements carried out in Figure 5.5-right, one can appreciate

that with the half-body plus one wing (with a mass of ∼ 50g), the requirements in power are

almost the half (1.9W ) to power flapping flight at 10Hz. In Figure 5.5-right, the input power

has been calculated as (Power= I2flapR), with Iflap = 341mA the electrical current required to

power the DC motor at the required torque.

The final flapping-wing mechanism is presented in Figure 5.6. Two gears rotate at maximum

625RPM to achieve the maximum wingbeat frequency of 10Hz. Transmission bars connect

each gear to the respective shoulder joint of each wing. The axe of gears rotation have been

placed to allow each wing to cover an angle of q1 = 120o during flapping (L1 = 2.5cm).

101

Page 134: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

gears

Shoulder

L1

L1

q1

z1

y1

x1

rigid Transmission bars

Figure 5.6: Detailed flapping-wing mechanism. Source: The author.

5.3.2.2 Membrane issues

In this stage (fixed-wing design), a first approximation to fabricate the wing-membrane was

tackled. In biological bats, the membrane (plagiopatagium) is an extraordinary tissue. It

contains dozens of tiny embedded muscles that control the tension and even tiny hairs on

the surface to sense airflow conditions and improve on flight control. These features are hardly

matchable with available artificial skin technology (polymers, etc). In this thesis, two features of

the plagiopatagium have been selected to replicate: i) stretchable and ii) cambering properties.

Figure 5.7 details the platform for the fabrication of the artificial membrane.

The membrane should be easy-expandable and not present a high load to the SMA actuators.

In our experiments, this property has been achieved by mixing 20g of two different compounds of

liquid soft platinum silicone rubber (http://www.smooth-on.com/Silicone-Rubber-an/c2_

1115_1129/index.html), cf. Figure 5.7-left, resulting in a light and stretchy wing membrane

with a thickness of 0.1mm. The fabrication platform consists on two surfaces of aluminium

with a thickness of 2cm. Both surfaces have the same dimensions and a mass of 1Kg-each.

The goal is having heavy surfaces that facilitate the distribution of the silicone mixture. The

resultant mixture is distributed between the surfaces and using four screw-system located at

the corners is possible to compress the mixture to achieve the distribution. Also, small pieces

of aluminium of 0.1mm-thick are located between the surfaces at each corner. This allows the

mixture of silicone has a thickness of 0.1mm when compressed by the screws.

By the end of the process, it is obtained a rectangular piece of silicone-membrane with a

thickness of 0.1mm, cf. Figure 5.7-right. However, hundreds of tiny air bubbles are formed

102

Page 135: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

aluminium surfaces

silicone rubber

siliconemixture

screws for separating the surfaces

air-bubbles

Figure 5.7: (left) platform for wing membrane fabrication, (right) membrane issues: air bubbles

are kept within the mixture. Source: The author.

during the compression process, which drastically affect the stretchable property. Because the

silicone-membrane is very thin, any uniformity within the structure causes the membrane to

fracture. To solve this problem, a vacuum chamber has been used. Previously to distribute the

silicone mixture between the surfaces, the solution is put in the vacuum chamber for about 10

minutes. This process eliminates the air bubbles, resulting in a smooth, flat and compact piece

of silicone-membrane. Section 5.3.5 will detail this achievement and assess the performance of

the membrane in terms of stretchable and cambering properties.

5.3.3 Step 4: Articulated-wing design

Having the design of the fixed-wing (step 3), this subsection presents the design of the artic-

ulated wing. The objective is to reproduce the most influential planar joints of the bat wing

skeleton. The model/CAD of this planar joints are detailed in Figure 4.2. As previously ex-

plained, biological bats have highly articulated wings, similar to the arms and hands of humans.

This level of wing dexterity enables fascinating control over the changing shape of the wings,

having this an impact on manoeuvrability. The designed wings have six degrees of freedom, but

only two are directly actuated: flapping (q1) and morphing (q3). The rest are under-actuated

joints that rotate via metal tendons. Figure 5.8 details the articulated-wing design. All param-

eters related to wing-morphology were previously covered in Chapters 3 and 4, cf. Figure 3.7

and Table 3.5.

Figures 5.8a-b show the metal tendons that enable the motion of the metacarpophalangeal-

(MCP) digits III, IV and V, which are attached to the wrist joint. Each digit has different

radii, allowing for different rotation ranges. Using this approach, the digits open and close

103

Page 136: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

Elbowjoint

Carpus/wristjoint

MCP-III digitshoulder

Supraspinatus

Scapula

Teres major

triceps

Humerus

spinous

ulna

Flexor carpi ulnaris

Radius

Biceps

extensor carpi radialis

IV digit

V digit

(a) (b)

(c) (d)

metal tendons

Figure 5.8: Articulated-wing design: a-b) metal tendons are placed inside the bones for con-

necting the elbow joint with the wrist. This enables each digit rotate as a function of the elbow’s

rotation, c) cartoon of a biological contracted wing and its parts, d) ABS fabricated articulated

wing. Source: The author.

to maintain the proper tension of the wing membrane during the morphing-wing motion. To

calculate the radii that form the joint of each digit and the key proportions to maintain during

wing contraction and extension, biological data was used. Figure 5.9b shows the angles and

proportions between digits that maintain the membrane tension. The data has been extracted

from in-vivo experiments depicted in Figure 5.9a. Without the membrane attached to the

skeleton, both elbow and wrist joints can rotate free of load, cf. Figure 5.9c, however, note

from the specimen’s pictures that the wings are never completely extended or retracted, and it

is precisely the membrane that constrains these maximum values for extension and contraction.

Once the artificial membrane is attached to the robot’s wings, cf. Figure 5.9d, it causes a load

that limits the range of the morphing-wings but maintain proper membrane tension aimed at

the proper generation of lift forces.

The following sections 5.3.4 and 5.3.5 details on the morphing-wing mechanism driven by

SMA-based artificial muscles acting as biceps/triceps and also on some properties of the silicone-

based membrane.

104

Page 137: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

WristIII

IV

V

WingtipMCPIII

15deg

~15deg

0.032m

0.014m

0.012m0.045m

35.4deg63.3deg

y

x

reference line

III

IV

V

Brown University

(a)

(b)

(c)

(d)

Figure 5.9: a) A high camera captures the motion of each marker places along the specime’s

wings. Pictures were taking during contraction and extension of the wings, b) detailed dimensions

of wing morphology during middle downstroke. Angles and proportions are extracted from the in-

vivo experiments from (a), c) membrane-free maximum rotation of elbow and wrist joints during

wing contraction and extension, d) maximum rotation of elbow and wrist joints during wing

contraction and extension including the wing-membrane load. Source: The author.

5.3.4 Step 5: Morphing-wing mechanism

Shape Memory alloys (SMA) power the morphing-wings. As previously mentioned along this

thesis, NiTi SMA wires have been selected because the closer relation with biological muscle

fibers: they can contract upon electrical heating and recover without any thermal process

involved. SMAs are also extremely light and small, being suitable for their use as artificial

muscles that compose the whole wig structure. In addition, SMAs have sensing capabilities,

105

Page 138: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

tendon-like chord attaching the muscle to the elbow

3lsma 2rj

SMA_1

SMA_2-+

Isma1 Fsma-1

Fsma-2

3

Isma2

SMA antagonistic model

4 mm of stroke

38.7 mm

SMA wires attached to metal strips

Figure 5.10: Antagonistic mechanism of SMA-based muscle actuators. Source: The author.

avoiding the need for external sensors that add additional weight and decrease the payload

capacity of the robot. Figure 5.10 details the antagonistic pair of SMA Migamotor artificial

muscles.

The SMA Migamotor actuators are connected to the elbow joint, enabling a range of motion

of ∼ 60o. During the downstroke, wings are fully extended in order to maximize the area and

increase lift, whereas during the upstroke, wings are folded in order to reduce aerodynamic

drag. In our robot, this property has been mimicked by attaching the antagonistic pair of SMA

actuators to the elbow, which allow the wing to track a reference trajectory by implementing

a proper control methodology (refer to Section 6.5 in Chapter 6). In Figure 5.10, the modules

SMA1 and SMA2 represent this antagonistic configuration (artificial biceps and triceps). When

one SMA actuator contracts, the generated pull force (Fsma) is transformed into a joint torque

τ3 (elbow’s torque). The input of each muscle corresponds to an electrical current signal,

denoted as Isma1 and Isma2 respectively, which are a direct function of the input heating power

Psma(t) = Isma(t)2Rsma(t), Rsma(t) being the SMA electrical resistance.

The mechanical design shown in Figure 5.10 is crucial for the hypothesis presented in this

thesis. It is precisely the capacity of changing the wing profile which enables efficiency in flight

control. Because the relevance of the morphing-wing mechanism and its direct implication

with the control approach, Chapter 6 details experiments that assess the response of the SMA

106

Page 139: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

migamotor muscles and the performance of the mechanism. Some of the experiments covered

are the follows:

• Wing modulation: it presents the closed-loop architecture that achieves proper SMA

actuation via PID control, cf. section 6.1 and 6.5.2.

• SMA Migamotor actuator characterisation: experiments for the identification of

morphing-wing actuation dynamics, cf. section 6.3.

• Accuracy of SMA actuator model: it asses the response of the identified model (the

plant) VS the real response of the SMA actuators, cf. section 6.3.2.

• SMA as sensors: it presents experiments that enable the relation between SMA resis-

tance changes and SMA strain (joint motion), cf. section 6.4.

• SMA performance: it presents experimental testing of SMA response in terms of speed,

output torque and accuracy, cf. section 6.7.2.

5.3.5 Step 6: The wing-membrane

Besides the SMA morphing-wing mechanical mechanism, the wing-membrane plays a key role

during the modulation process of the wing. This section complements section 5.3.2.2 with details

regarding the testing of the silicone membrane in terms of: i) stretchable and ii) cambering

properties.

stretchable

Figure 5.11 shows empirical testing regarding the stretchable property of the artificial mem-

brane. To achieve this property, Dragon Skin Series silicones with high performance platinum

cure silicone rubbers were used. Both silicone components were mixed 1A:1B by weight or

volume and cure at room temperature with negligible shrinkage. Cured Dragon Skin is very

strong and very stretchy. It will stretch many times its original size without tearing and will

rebound to its original form without distortion. As previously highlighted, vacuum degassing is

recommended to minimize air bubbles in cured rubber. Figure 5.12 details a technical overview

of Dragon Skin silicone properties. To fabricate a thick membrane surface, it is recommended

to mix the three components of Dragon Skin, e.g. 10-medium, 20 and 30. On the other hand,

whether the objective is to fabricate a thin membrane surface, only two components are rec-

ommended, e.g e.g. 10-slow and Dragon Skin 20.

107

Page 140: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

(a)

(b) (c)

Figure 5.11: Testing on the stretchable property of the fabricated silicone membrane: a) ex-

tended plagiopatagium skin, embedded tiny muscles control the membrane tension, b) the fabri-

cated silicone membrane is attached to the wing skeleton using Sil-Poxy RTV, c) the fabrication

process described in section 5.3.2.2 results on a light artificial skin with the required stretchable

property. Also note that after vacuum, air-bubbles are eliminated. Source: The author.

For preparation, it is recommended to use a properly ventilated area (room size ventilation).

Wear safety glasses, long sleeves and rubber gloves to minimize contamination risk. Wear vinyl

gloves only. Latex gloves will inhibit the cure of the rubber. Store and use material at room

temperature (73F/23C). Warmer temperatures will drastically reduce working time and cure

time. Storing material at warmer temperatures will also reduce the usable shelf life of unused

material. These products have a limited shelf life and should be used as soon as possible. Before

mixing the components, pre-mix one thoroughly. After dispensing required amounts of parts

A and B into mixing container (1A:1B by volume or weight), mix thoroughly for 3 minutes

making sure to scrape the sides and bottom of the mixing container several times. After mixing

parts A and B, vacuum degassing is recommended to eliminate any entrapped air. Vacuum

material for 2-3 minutes (29 inches of mercury), making sure that you leave enough room in

container for product volume expansion. It is not recommended to exceed 10% by weight of

total system (A+B).

108

Page 141: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

Figure 5.12: Technical overview of Dragon Skin silicone properties. Source: http://www.

smooth-on.com.

Figure 5.13: Mid-downstroke wing camber and angle of attack are estimated as follows: (A) A

parasagittal (xg-zg) cross section of the wing was taken at the yg-value of the wrist at the time

of maximum wingspan. Six triangular sections of the wing membrane crossed that plane and the

intersections of triangle borders in the plane (red circles) were used as estimates of membrane

position. (B) The actual curved shape of the membrane in the plane (solid black line) was

estimated using the first term of a sine series fitted to those seven points. The maximum distance

of the membrane line from the chord line (dashed grey line) was divided by the length of the chord

line to give wing camber. (C) Angle of attack (α) was calculated as α1+α2, where α1 is the angle

of the wing chord line above horizontal (blue dashed line), and α2 is the angle between horizontal

and the velocity vector of the wrist (red arrow) in the xg − zg plane. Source: (20).

Cambering

In (20), wing camber has been quantified for several species of biological bats. The results of

the study have shown the influence of bat body mass (mt) not only on the wing camber, but

also on the lift production, flight velocity, and flapping frequency. Figure 5J in (20) shows the

wing camber coefficient as a function of body mass. For the specie Rousettus aegyptiacus, wing

109

Page 142: BaTboT: a biologically inspired flapping and morphing bat robot

5.3 BaTboT mechanics

camber at maximum span is proportional to ∝ m0.71t . Figure 5.13 has been extracted from

(20) and explains the procedure to measure the mid-downstroke wing camber using an external

camera. Following the relationm0.71t , beingmt = 0.079kg the mass of BaTboT without battery,

wing camber has been empirically adjusted to be about 0.16, similar to the values measured

from the biological counterpart (0.14 − 0.15). Figure 5.14 shows wing camber values. The

variation of the maximum camber of the membrane airfoil shape with respect to time is shown

for a representative case (Vair = 4.1ms−1). Over a significant portion of the wingbeat cycle

(arguably the portion during which most of the aerodynamic forces are generated by the wing),

the camber is relatively constant and hovers around the value of 10% of the chord length. The

spikes in the value of the camber near the top of the upstroke are to be expected since wing

will fold significantly during this portion of the wingbeat.

Figure 5.14: Variation of camber during the wingbeat cycle. Source: (25).

On the other hand, we have performed aerodynamics measurements in the wind-tunnel for

testing how the membrane behaves when subject to different airflow speeds. These experimental

results are reported in Figure 5.15. The robot is mounted in the wind tunnel, on the end of

a supporting sting that defines the angle of attack. The robot is mounted on top of a 6-DoF

force sensor from which both lift CL and drag CD coefficients are experimentally calculated as

a function of the airflow speeds and angle of attack. Typical results are shown in Figure 5.15.

In Figure 5.15a the membrane is fixed to a support base and camber is not adjusted. Because

of the airflow (Vair = 5ms−1) at an angle of attack varying from 10− 20o, an excessive camber

value is measured with the camera (0.48). Lift and drag coefficients (CL, CD) are measured

between the range of interest (angle of attack varying from 7 − 13o). On the other hand, in

Figure 5.15b, the same membrane surface is attached to BaTboT wing’s skeleton and camber

is adjusted to be about 0.16. Note the robot is not flapping, the wings are fixed to middle

110

Page 143: BaTboT: a biologically inspired flapping and morphing bat robot

5.4 BaTboT electronics and sensors

(b)

Force sensor

airflow

FL

FD

excesivecambering

load cell

CL

CD

airflow propercambering

(a)

(b)

CL

CD

Figure 5.15: Implications of wing camber into lift and drag production (Vair = 5ms−1, fixed-

wing): a) excessive wing camber (0.48), b) proper wing camber (0.16). Source: The author.

downtroke. Lift and drag coefficients are again measured under the same airspeed conditions

and note there has been an increment of 65% in lift coefficient value (at angle of attack of 9o).

Further experiments in Chapter 6, cf. Figure 6.4 shows quantification of the membrane

loads caused during the extension process of the wings.

5.4 BaTboT electronics and sensors

BaTboT is composed by a simple electronic configuration mainly because each added com-

ponent represents an additional weight to the overall structure. The challenge is to design an

onboard hardware architecture light and power enough to enable the proper control of the robot

towards its autonomous operation. This section details the onboard hardware architecture, its

components, functions, and the power consumption of the robot.

5.4.1 Arduino controller-board

Arduino is one of the most extended, simple and robust commercial solutions of micro-controllers.

The ArduinoNano version (http://arduino.cc/en/Main/ArduinoBoardNano) is one of the

lightest chips powered by an Atmel ATmega168 running at 16MHz. It has an operation volt-

111

Page 144: BaTboT: a biologically inspired flapping and morphing bat robot

5.4 BaTboT electronics and sensors

+

Figure 5.16: Micro-controller and Migamotor SMA muscle connection diagram. Source: The

author.

age of 5V , 14 I/O digital pins (6 are PWM) and 8 analog pins. The ATmega168 has 16KB of

flash memory for storing code (of which 2KB is used for the bootloader). Arduino provides a

software (http://arduino.cc/en/Main/Software) for programming the micro-controller. An-

nex 10.5 details the control-code and Kalman filter code programmed into the Arduino Nano

for driving BaTboT behaviour, whereas Figure 5.16 shows the connection diagram between the

micro-controller and the Migamotor SMA actuator.

The SMA actuator can provide simple full-range motion by simply having their CNTRL pin

enabled. The Migamotor SMA muscle will keep contracting/rotating as long as the CNTRL

pin is held high until it reaches its full extent. The rate at which this motion will occur will vary

depending on, the load, friction in any linkage, voltage level and ambient temperature. The

recommended way of controlling speed is using a Pulse Width Modulation (PWM) signal to the

CNTRL pin rather than simply setting it. The speed of motion will be governed by the duty

cycle of this PWM signal. The larger the duty cycle, the faster the Migamotor SMA muscle will

move. Within certain constraints the actual period of the PWM does not have any affect. Unlike

the control of some other devices, Migamotor SMA muscle actuator can operate at relatively

low periods (typically to between 100Hz and 1KHz). The Migamotor SMA muscle actuator

has a position feedback pin that gives an analog signal that can be interpreted to provide an

approximate position. The change that must be measured is the change in resistance between

the position feedback pin and the Vsma pin of the Migamotor SMA muscle. This resistance will

112

Page 145: BaTboT: a biologically inspired flapping and morphing bat robot

5.4 BaTboT electronics and sensors

vary around 10% during the contraction of the Migamotor SMA muscle. There are a number

of methods that can be used to measure this resistance. A typical circuit, cf. Figure 5.16 used

to measure this resistance is to time a capacitor charge up that is inserted in series with the

resistance of the position control pin.

In Figure 5.16, setting the CNTRL signal high of the Migamotor SMA muscle will drive the

position feedback pin low and cause the capacitor to discharge. Clearing the CNTRL signal

will start the capacitor charging. In order to be able to accurately time this charge up period,

the circuit uses a comparator with a Voltage Reference input that will trigger an interrupt in a

micro-controller when the reference voltage is reached. The code required for a resistance based

position control system configured around this is very similar to that of the optical encoder.

The only major difference is that instead of reading pulses from the encoder in an interrupt

handler, the feedback resistance using the capacitor circuit must be periodically measured. The

pseudo code for such a reading is detailed as follows:

1. Set Voltage Reference to low value (e.g. 0.8v)

2. Set Migamotor SMA muscle CNTRL Pin High to discharge capacitor

3. Wait for Comparator to indicate Voltage has dropped below VRef

4. Set Voltage Reference to mid value (e.g. 1.5v)

5. Set Migamotor SMA muscle CNTRL Pin Low to start capacitor charging Wait for Comparator to indicate

Voltage has risen above VRef Start Timer

6. Set Voltage Reference to high value (e.g. 3.5v)

7. Wait for Comparator to indicate Voltage has risen above VRef Stop Timer

8. Resistance = Timer Value

By measuring the SMA resistance changes is possible to estimate the position state driven

by the actuator and therefore to close the control loop. This SMA Resistance-to-Motion rela-

tionship (RM) is introduced in Chapter 6 section 6.4, whereas the control algorithm is detailed

in section 6.5.4.

5.4.2 The Inertial Measurement Unit (IMU)

The Inertial Measurement Unit is a digital combo board of 6 Degrees of Freedom with gyros:

ITG3200 and accelerometers: ADXL345 (https://www.sparkfun.com/products/10121?). IMU

readings are vital for the feedback of attitude variables that enable roll and pitch control of

BaTboT. Details about the use of the IMU merged with the control architecture can be found

in Chapter 6, Figure 6.14. This chip is ready to be connected to the Arduino board, the

unique requirement is to filter the IMU data for reducing noise. This procedure is shown in

113

Page 146: BaTboT: a biologically inspired flapping and morphing bat robot

5.4 BaTboT electronics and sensors

roll IMUroll IMU filtered

pitch IMUpitch IMU filtered

sample

attit

ude

[deg

]

Figure 5.17: Comparison of attitude IMU readings before and after Kalman filtering. Source:

Arduino.

Figure 5.17. Kalman filtering technique is used for both reducing noise and predicting attitude

motions. Annex 10.5 details the code.

5.4.3 SMA power drivers

The Miga Analog Driver V5 (MAD-V5) is a MOSFET switch designed to safely power the Mig-

amotor SMA actuators across a wide range of speeds or input voltages. This driver generates

the current signal based on the digital control command sent from the Arduino board. The

schematic of the circuit is described in Figure 5.18. The MAD-V5 allows either push-button

operation, or external GATE (CNTL) signals to actuate the Migamotor SMA actuator until

the END limit is reached (goes LOW). The MAD-V5 then cuts power momentarily, preventing

overheating of the SMA wires. The Gate transistor allows up to 30V input, but it is rec-

ommended to use logic (2.5 to 5-Volt) levels. Pulse-Width-Modulated (PWM) signals can be

applied at the Gate to control the actuation speed for a set voltage, or even an AC driven

current signal mounted on a DC level. For instance, the application of +28VDC power to the

Migamotor SMA actuator would result in very fast actuation (∼ 70ms). In Figure 5.18, JP1 is

the power supply and/or micro-controller connector, JP2 is connected to the Migamotor SMA

actuator. Maximum peak current: 7A, maximum continuous current: 5A.

114

Page 147: BaTboT: a biologically inspired flapping and morphing bat robot

5.5 BaTboT consumption

Control signal

SMA-mosfet power

1.5cm

Figure 5.18: Miga analog driver V5 pinout diagram. Source: The author.

5.5 BaTboT consumption

Consumption is measured in terms of required electrical currents. Figure 5.19 shows the per-

centage of current consumption of each electronics component of BaTboT, whereas Table 5.5

consigns the numerical values. As expected, the SMA actuators require most of the input

current, about 78.7%.

Table 5.5: General values of current consumption

Component Required current [A] % of consumption

SMA actuators (x4) 0.6x4 78.7%

DC motor (x1) 0.35 13%

Arduino board (x1) 0.05 3%

IMU (x1) 0.1 5%

Total consumption = 2.9

115

Page 148: BaTboT: a biologically inspired flapping and morphing bat robot

5.6 BaTboT costs

Figure 5.19: Percentage of current consumption per component. Source: The author.

5.6 BaTboT costs

Table 5.6 details the costs of the components that are involved in the fabrication process of the

robot.

Table 5.6: Fabrication costs

Component Manufacturer/seller Item cost Quantity used Total

SMA Nanomuscle NM70R-6P Migamotor 6.93e 4 27.72e

Miga Analog Driver V5 (MOSFET) Migamotor 10.75e 4 43e

DC adjustable conversor (4-25V-2A) BricoGeek 10.40e 1 10.40e

50:1 Micro Metal Gearmotor HP Pololu 12.30e 1 12.30e

IMU 6DOF (ITG3200/ADXL345) Sparkfun 50e 1 50e

ArduinoNano V3.0 Arduino 17.95e 1 17.95

Polymer Lithium Ion Battery - 850mAh Sparkfun 6.90e 1 6.90e

(ABS) plastic material - - - 60e

Aluminium plates for membrane molde alu-stock 40e 2 80e

Steel wire rope for metal tendons mcmaster 5e 1 5e

Sil-Poxy Silicone Adhesive smooth-on 22.13e 1 22.13e

Dragon Skin Silicone Rubber for wing membrane smooth-on 23.20e 2 46.40e

Total costs =381.8e

5.7 Remarks

This Chapter has completely described the bio-inspired design-flow applied to the development

and fabrication of BaTboT. The criteria for design have been classified in terms of morphology,

kinematics, dynamics and aerodynamics parameters extracted from the analysis of biological

data. Most importantly, the data has been related as a function of design parameters such as

overall mass, wingspan, body mass, wing mass, wing lenghts, etc. These relations have enabled

116

Page 149: BaTboT: a biologically inspired flapping and morphing bat robot

5.7 Remarks

the definition of a framework that can be applied to future designs of BaTboT. The Chapter has

done an special emphasis into the design of the highly articulated morphing-wings, showing how

to mechanically achieve the wing modulation required for testing the hypothesis proposed at

the beginning of this thesis. The following Chapter introduces the characterisation of the main

components introduced herein and the control methods that are key for the proper regulation

of the wing profile. All of this towards future developments that achieve autonomous flight.

117

Page 150: BaTboT: a biologically inspired flapping and morphing bat robot

6

BaTboT Control

6.1 Control goal

The control goal is twofold: firstly, enabling BaTboT of accurate forward and turning flight

capacity, and secondly achieving efficient flight by properly driving wing modulation. First of

all, the control scheme must ensure the robot is able to track wing joint trajectories defined in

Figure 4.5 (forward flight), and Figure 4.6 (turning flight). In case of forward flight maneuver,

when both wings are flapping symmetrically and keeping the same shape-configuration, the

resultant torques produced at the robot Center of Mass (CM) are theoretically cancelled between

each other. In the practice though, a slight mechanical bias caused within the fabrication

process introduces asymmetries that only could be regulated by means of feeding back attitude

measurements (roll and pitch) and adjusting the attitude tracking error. In case of turning

flight maneuver, wings flap in different shape-configurations aimed at displacing the CM and

therefore causing the attitude variation of the body. The control must monitor the center of

mass is on the desired place.

This chapter introduces the approaches-to-follow to fulfill both goals. To this purpose a

Flight Control Architecture (FCA) is formulated in Figure 6.1. It is composed by two control

strategies:

1. Morphing-wing controller: it consists on an inner closed-loop position control which drives

the changes on wing configuration via Shape Memory Alloy actuation. Wings are mod-

ulated depending on the wing joint references trajectories and an outer attitude control

loop commands. The control goal is to efficiently drive Shape Memory Alloy ac-

tuators aimed at the fastest modulation of the wing shape, i.e, contraction and

118

Page 151: BaTboT: a biologically inspired flapping and morphing bat robot

6.2 Flight Control Architecture (FCA)

SMAactuation

Rsma=Vsma/Isma

SMA Resistancemeasurement

Attitude control

Motors

-+ Morphing

control

outer loop

q3

R/C (flapping)IMU

Bat robot

+-

R/C (attitude)forward/turning

inner loop

[Uheating]

[q1…q6]R,L[roll, pitch]

q3 = f (R)

q1,ref

φ,θ[ ]

uφ ,uθ⎡⎣ ⎤⎦

wristmapping

Elbow motionestimation

uθ = q2,ref

uφ = q3,ref

Wing trajectories

τ 3

τ1τ 2

τ 3τ 4τ 5τ 6

Figure 6.1: Flight Control Architecture (FCA). Source: The author.

extension. Efficiency is evaluated as a tradeoff between SMA output torque

and power consumption that achieve the fastest SMA actuation speed.

2. Attitude controller: It drives the inner control loop. It consists on an outer closed-loop

position scheme based on a novel backstepping+DAF controller. It regulates the roll

(φ) and pitch (θ) response of the robot and allows forward and turning maneuvers. The

control goal is to consider the influence of wing inertia into the attitude control

law aimed at efficiently driving the wing modulation. Efficiency is evaluated

in terms of thrust production.

6.2 Flight Control Architecture (FCA)

Inner loop: morphing-wing modulation

It is composed by six modules:

• Motors: it refers to the DC-motor models that power the primary flapping motion q1

and the rotation of wings forward/backward about the gravity axis q2. The latter allows

for the pitch motion of the body θ. Outputs are torque commands to the inertial model;

τ1, τ2.

• SMA actuation: it refers to the Shape Memory Alloy antagonistic model. SMA actuators

are driven by the control signal uheating which corresponds to the input heating power.

119

Page 152: BaTboT: a biologically inspired flapping and morphing bat robot

6.2 Flight Control Architecture (FCA)

Each SMA contracts upon electrical heating generating a pulling force on the elbow joint

of the wing. The output is the joint torque that rotate the elbow joint: τ3.

• Wrist mapping : it maps joint torques τ4, τ5, τ6 from elbow torque τ3. This is due to wrist

jointsq4, q5, q6 are under-actuated. The mapping is based on the kinematics configuration

of the wrist joint shown in Figure 3.7b.

• Morphing-wing : it refers to the control strategy that drives SMA actuators. In this

chapter two control techniques are evaluated: i) a sliding-model nonlinear control and ii)

a PID linear control. Inputs are the position errors of elbow joint motions of both wings

(q3,ref − q3)R,L and outputs are the control signals uheating that drive SMA actuation.

• Elbow motion estimation: it refers to the Resistance-Motion relationship between SMA

electrical resistance Rsma and joint motion q3. Besides actuators, SMAs can also be used

as sensors. It turns out that Rsma changes lineally with respect to the strain rate of

the SMA wire. Based on this, elbow motion can be estimated as a function of electrical

resistance measurements: q3 = f(Rsma).

• SMA resistance measurements : it refers to the calculation of SMA electrical resistance

Rsma by measuring the applied voltage and electrical current through the SMA wires.

The inner loop runs at 30Hz. It allows a maximum SMA physical actuation of 2.5Hz

mainly due to the limitations of SMA actuation caused by slow cooling time of the wires. This

upper limit in actuation speed was previously estimated by simulations from Figure 4.22a, cf.

Table 4.9. Experiments carried out in section 6.3 confirm this limit value using the real system.

Further details of the inner loop control can be found in section 6.5.

Outer loop: attitude regulation

It is composed by three modules:

• Wing trajectories : it refers to joint trajectory generator module. It generates kinematics

profiles q1...q6 for each wing depending on the maneuver type: forward or turning flight.

Profiles are those previously described in Figures 4.5 and 4.6 respectively.

• Attitude control : it refers to the control strategy that drives the attitude regulation (φ, θ)

and manages the inner loop. Inputs are the desired joint trajectories and attitude ref-

erences. Outputs are control signals that drive the inner loop, allowing for proper wing

modulation. The attitude control is based on a novel backstepping+DAF methodology,

120

Page 153: BaTboT: a biologically inspired flapping and morphing bat robot

6.2 Flight Control Architecture (FCA)

which uses a Desired Angular Acceleration Function (DAF) to incorporate the influence

of wing inertia into the control law.

• IMU : it refers to the Inertial Measurement Unit sensor, which allows feeding back attitude

variations. In the practice, the IMU sensor includes 3-axis accelerometer and 3-axis

gyroscopes1.

The outer backstepping+DAF attitude control runs at 20Hz and generates proper control

references that allow the wing modulation.

The methods for control

1. Experiments on frequency response analysis are carried out for the identification of the

SMA actuation module, cf. Figure 6.1. System identification is aimed at characterizing

the relationship between applied power input (via electrical current heating) and SMA

output torque. Section 6.3 details the experimental methods.

2. Experiments are carried out to characterize the Resistance-Motion (RM) relationship

between SMA electrical resistance changes (Rsma) and SMA strain. This allows the

identification of the Elbow motion estimation module in Figure 6.1. The RM relationship

makes possible to feedback elbow motion q3 and close the inner loop. Section 6.4 details

the methods.

3. The Morphing wing control module is formulated (inner loop). Two control methodologies

are first tested on simulation, one based on sliding-mode technique and the other based

on classic PID control. Also, mechanisms that enhance the performance of the SMA

actuators are complemented into the morphing controller. These mechanisms are called

anti-slack and anti-overload. Section 6.5 details the methods.

4. The Attitude control module is formulated (outer loop). It shows how the properly reg-

ulate roll and pitch variables (φ, θ) based on wing modulation. A novel strategy called

backstepping+DAF is introduced. Using the inertial model from Eq. (4.14) and (4.15), a

Desired Angular Acceleration Function (DAF) is formulated. The DAF uses the inertial

model to incorporate wing inertia information that is key for the generation of body accel-

erations. These accelerations are used as references to the backstopping+DAF controller

1The IMU-model placed on the robot corresponds to ITG3200/ADXL345, http://www.sparkfun.com/

products/10121.

121

Page 154: BaTboT: a biologically inspired flapping and morphing bat robot

6.3 SMA actuation: experimental characterization

aimed at improving the modulation patterns of wing kinematics. Section 6.6 details this

novel control.

6.3 SMA actuation: experimental characterization

6.3.1 Frequency analysis

Previously in Chapter 4 section 4.5.2 a phenomenological model was used to describe the com-

plex nature of SMAs, especially in terms of their thermomechanical behaviour and the hys-

teretic effects. This becomes useful for analyzing over-heating effects and other aspects such

as over-stressing, that could be impractical through real experimentation. However, to control

purposes, phenomenological models are mostly useless. The main reason concerns that some

of the internal state variables of the SMAs are irrelevant to control, such as temperature or

martensite ratio. Actual measurements of these variables are often difficult, even impractical.

On top of this, phenomenological models describe the large-signal behavior of SMAs, which is

quite nonlinear, hysteretic and often not repeatable.

To overcome these problems, remarkable investigations in (9) and (67) about the frequency

response analysis of NiTi SMA actuators were carried out aimed at the development of linear

force models that attempted to characterize the relationship between applied power and SMA

output force. This relationship is actually more suitable than phenomenological models for the

development of controllers that drive SMA actuation based on electrical heating. The procedure

described in (9) demonstrated that the AC frequency response of SMA NiTi wires is similar

to that of a first-order linear system.This leads to the analyzes of SMA small-signal behavior,

rather than the large-signal response.

This section briefly describes the methods for obtaining a SMA force model relating output

force -to- input heating power. As demonstrated in (9), the frequency response analysis allows

for the study of the small-signal response of SMA wires over a suitable frequency range, which

has been observed to be very repeatable and also exhibits very little hysteresis. So, the the goal

here consists on finding the linear model plant that matches the behavior of the Migamotor’s

SMA nano-muscles in terms of input power to output torque. Frequency response analysis is

carried out by applying sinusoids test signals to the input acting as heating power, and looking

at the output to see how the system responds by measuring the output force. Therefore, the

model to be identified can be represented by a transfer function that matches the measured

gain and phase data over the frequency range of interest. Here, the input signal command

122

Page 155: BaTboT: a biologically inspired flapping and morphing bat robot

6.3 SMA actuation: experimental characterization

IsmaDAC

16-bit ADC + anti-aliasing filter9620-05-ATI

Fixed wing testbed

Low passfilter

Torqueconversion

Fτ3

Currentconversion

uheating Transconductanceamplifier

lr

q3~60º

F

Force sensor

τ3

SMAs

Figure 6.2: Experimental testbed for the characterization of SMA input power (uheating) to out-

put torque (τ3). Forces (F ) are measured using a force sensor with 0.318 gram-force of resolution.

Torque conversion is applied by considering the humerus bone length (lr), as: τ3 = Flr[Nm]. This

allows the identification of SMA actuation. Source: The author.

represents the heating power, which can be described as: uheating = a+ bsin(ωt). The term a

is the mean input power, b is the magnitude of the sinusoidal component, and ω = 2πf , being

f the commanded frequency for SMA contraction.

Experimental Testbed

Figure 6.2 details the experimental setup for the identification of the SMA actuation. Applied

heating power uheating is converted to a current signal using a nominal value for the SMA wires

electrical resistance provided by the manufacturer Rsma = 8.5Ω (cf. www.migamotors.com),

and is then converted to an analog voltage and sent to the precision transconductance amplifier

that drives the SMA actuator. The force response is measured using a Nano 17 transducer

ATI Industrial Automation force sensor with0.318 gram-force of resolution. In addition, a

16-bit DAC with embedded 1st order anti-aliasing filter with a bandwidth from 5KHz to

10KHz is also used. Details about the specifications of the force sensor can be found at:

http://www.ati-ia.com/products/ft/ft_models.aspx?id=Nano17.

The force signal F is recorded into the computer, and then mapped to a torque value τ3.

This corresponds to the elbow torque. Note the force measurements have been carried out

taking into consideration the payload of the bones and the friction of wing joints1, but the

membrane load has not been taken into account. Later in section 6.3.2 it will be demonstrated

that thanks to the highly stretchable and light properties of the membrane, the resultant data

from the identification process is not significantly affected.

1Joint friction is not directly measured or quantified.

123

Page 156: BaTboT: a biologically inspired flapping and morphing bat robot

6.3 SMA actuation: experimental characterization

The identification procedure

To identify the system, the magnitude and phase of the sine-wave component of F that is

generated upon heating driven by uheating = a+bsin(ωt) is measured over a frequency range of

0.1 to 100Hz. It is important to highlight that the DC component of uheating (a) should be keep

constant over the frequency range. To extract the sine-wave component in the recorded force

data F a least-squares methodology is used. The least-squares-based parameter identification

algorithm is an extended methodology widely applied in several systems. A detailed step-by-

step use of this method can be found for a similar SMA application in (9). Here, a briefly

introduction of the least-squares methodology is presented. Considering the measured force

data F (t) of the SMA is of the form:

F (t) = A cos (ωt+ ϕ)F (t) = A [cos (ϕ) cos (ωt)− sin (ϕ) sin (ωt)] ,

(6.1)

only two variables are known: i) the measured force data F (t), which is produced when the

SMAs are pulling the elbow joint when contract upon heating (uheating), and ii) the driving

frequency ω = 2πf , being f the actuation frequency of uheating. However, the force amplitude

A, and the phase of the signal ϕ are unknown. The goal is to estimate both parameters.

In order to put together the unknown parameters, a parametric model is used in (6.2), being

z (t) = F (t). Note both unknown parameters A, ϕ are now contained into vector θ∗.

z (t) = θ∗Tφ(t)z (t) = [A cos (ϕ) , A sin (ϕ)]

T[cos (ωt) ,− sin (ωt)]

T (6.2)

By using the Adaptive Control Toolbox in ©Matlab, a least-squares algorithm has been

used to estimate: θT = G(t)e(t), where G(t) is a gain vector and e(t) corresponds to the

estimation error: e(t) = z(t) − z (t). The term z (t) corresponds to the parametric estimation

model z (t) = θTφ(t). In other words, the error reflects the distance between the estimate θ

and the unknown θ∗. The least-squares algorithm fits a mathematical model to a sequence of

observed data by minimizing the sum of the squares of the difference between the observed and

the computed data. Once θ∗ has been estimated as θ = [A1, A2]T , being A1 = Acos(ϕ), and

A2 = Asin(ϕ), the estimation of A and ϕ of force amplitude A and phase ϕ is defined as:

A =√A2

1 + A22,

ϕ = tan−1(A2

A1

) (6.3)

124

Page 157: BaTboT: a biologically inspired flapping and morphing bat robot

6.3 SMA actuation: experimental characterization

Mag

nitud

e (d

B)Ph

ase

(deg

)

Bode Diagram

Frequency (rad/s)

Experimental

Frequency [rad/s]

Experimental

Model

Model

Phas

e [d

eg]

Mag

nitu

de [d

B]

Figure 6.3: (Experimental VS model) Bode magnitude and phase plots for NiTi 150μm SMA

Migamotor actuators. The insets show several experimental measurements of magnitude and

phase. Magnitude is given by: 20log(A/b), where A is the least-squares estimation of the force

amplitude measured using Eq. 6.3, and b is the AC power of the input signal uheating. Phase is

given by the term ϕ in Eq. 6.3. The transfer function that fits the experimental data is shown

in Eq. 6.4. This plot also compares the model in Eq. 6.4 against the experimental data. Source:

The author.

After running the identification algorithm, Figure 6.3 shows the experimental Bode plots.

The magnitude of the curves is given by 20log(A/b), where A is the least-squares estimation of

the force amplitude, and b is the AC power of the input signal Uheating. Note the SMA frequency

response in Figure 6.3 is similar to that of a first-order linear system, where the frequency at

which the change of slope in magnitude occurs is known as a pole (s = −2.857). Several

measurements have indicated that the suitable transfer function that fits the experimental data

is:

τ3(s) = 0.016(0.35s+ 1)−1uheating(s) (6.4)

By using the model in Eq. (6.4), the AC behavior of SMA actuation has been experimentally

identified in the form of a Laplace transfer function that relates the input heating power uheating

125

Page 158: BaTboT: a biologically inspired flapping and morphing bat robot

6.3 SMA actuation: experimental characterization

with the SMA output torque generated at the elbow joint τ3. The following section explores

the accuracy of the identified SMA actuation model to control purposes.

6.3.2 Experimental validation of SMA actuation model

Using the setup from Figure 6.2, this section introduces a comprehensive experimental review

on the accuracy of the identified SMA actuation model in Eq. (6.4). Three experiments are

carried out:

1. The membrane-load : The fabricated silicon-based membrane produces an external load to

the SMA actuator. Despite this load is minimum due to the light and stretchable property

of the membrane material, it seems convenient to quantify and verify the capacity of the

SMA actuators interns of force production regarding the membrane load. Quantification

is achieved by measurements of output torque (τ3) and required input power (uheating).

2. Morphing-wing frequency : SMA actuators allow the wings to contract and extend via

elbow and wrist rotations. This is known as morphing-wings. Here, nominal and over-

loaded operation modes of the SMA actuators are evaluated in terms of actuation speed.

Actuation speed is experimentally calculated by measuring the rotation speed of the el-

bow joint (θ3). Also, experimental data is compared against simulations with the SMA

phenomenological model (cf. Figure 4.22a and Table 4.9).

3. SMA actuation model accuracy : It compares the nominal and overloaded response of the

model in Eq. (6.4) against experiments of small-signal response of the SMA actuators at

driven frequency of f = 2Hz.

The membrane-load

Figure 6.4 shows experimental measurements of generated SMA output torque at elbow joint

(τ3) in response of nominal and overloaded values of input heating power (uheating). The

applied heating power corresponds to electrical current bias values ranging from Isma = 300mA

(nominal) to Isma = 550mA (overloaded). Theoretically, Figure 4.22a in Chapter 4 defined that

in order to achieve the required rotation speed of the elbow joint (θ3) under both nominal and

overloaded modes of SMA operation, DC input currents ranging from I = 350mA to I = 550mA

should be applied. Simulations indicated the elbow joint is able to rotate ∼ 60o in 200ms

when applying a nominal power of uheating =∼ 1.04W , and by increasing the applied power

to uheating =∼ 2.57W :∼ 60o in 100ms (overloaded). In the practice thought, joint frictions

126

Page 159: BaTboT: a biologically inspired flapping and morphing bat robot

6.3 SMA actuation: experimental characterization

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.0180

0.5

1

1.5

2

2.5

3

3 [Nm]

Powe

r[W]

Measurement Measurement

u heatingW [

]

τ3τ3

Figure 6.4: (Above) right wing extended and contracted taking into account the load produced

by the silicon membrane. (Below) Experimental measurements of SMA output torque τ3 generated

by input heating power uheating. Values are classified by nominal and overloaded operation mode

of the SMA actuators. Nominal behavior is achieved by applying uheating =∼ 1.36W whereas

overloaded behavior requires an input power of uheating =∼ 2.57W . Inset plots show the average

peak of produced elbow torque at both SMA operation modes. Source: The author.

and the anisotropic loading of the membrane make the input power to increase with respect

to the simulation predictions. In the experiments, under nominal mode (uheating =∼ 1.36W )

the output torque τ3 stabilizes around ∼ 0.007 − 0.008Nm whereas under overloaded mode

(uheating = 3.06W ) around ∼ 0.016−0.02Nm. It has been observed in Figure 6.6 about 20% of

increase in the input heating power to produce the desired output torque that achieves elbow

joint rotation speeds shown in Figure 6.5. It is precisely the high stretchable property of the

membrane material that allows for minimum load. The following section emphases in this issue.

Also, numerical values comparing simulation and experimental data can be found in Table 6.1.

Morphing-wing frequency

Figure 6.5 presents the average results of several elbow joint rotation speeds (θ3) achieved af-

ter applying both nominal and overloaded profiles of input power uheating. Root mean square

127

Page 160: BaTboT: a biologically inspired flapping and morphing bat robot

6.3 SMA actuation: experimental characterization

0 1 2 3 4 5 60

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Measurement

3[deg

/ms]

OverloadedNominal

0 1 2 3 4 5 6Measurements

0

q 3deg/ms

[]

0.10.20.30.40.50.60.7

0

Figure 6.5: Experimental measurements of joint rotation speeds of elbow joint (θ3) that are

obtained by applying heating power values at nominal (uheating = 1.36W ), and overloaded

(uheating = 3.06W ) SMA operation. Source: The author.

(RMS) joint-speed values have been measured using the -contracted - and -extended - pins pro-

vided by the SMA artificial muscles (cf. section 5.3.4). These pins activate when the actuator

reaches the maximum stroke (contracted) and when it returns to the neutral position (ex-

tended). In the nominal mode of SMA operation, the joint speed average is ∼ 60o in 225ms

(0.26deg/ms), whereas in the overloaded mode, ∼ 60o in 120ms (0.5deg/ms). Thereby, regu-

lating SMA operation within the overloaded region enables the wings to contract and extend

at a maximum frequency of f = 2.5Hz. This is known as the morphing-wing frequency. This

value is obtained as follows: (tdown + tup + tcool = 400ms)−1, where tdown and tup are the

times involve in the downstroke and upstroke cycles of the wing (max. speed of each cycle

set at 120ms) and tcool = 160ms is the death time required to enable the active SMA wire to

recover its original length during the cooling process. The death time is important within the

antagonistic configuration of SMA actuators since it contributes avoiding lock issues. This will

be explore later in section 6.5.

SMA actuation model accuracy

The nonlinear SMA phenomenological model described in Chapter 4 has been useful to ad-

dress some important limitations of SMA thermo-mechanics, at least in terms of simulation.

Previously in Eq. (6.4), a frequency response analysis allowed for the identification of the AC

behavior of the SMA actuators. To verify the accuracy of the model upon small-signal heating

128

Page 161: BaTboT: a biologically inspired flapping and morphing bat robot

6.3 SMA actuation: experimental characterization

0 0.5 1 1.5 2 2.5 3 3.5 4 4.50.002

0.004

0.006

0.008

0.01

0.012

0.014

0.016

0.018

0.02

0.022

3 [Nm]

t[s]

uheating=3.06W(overloaded)

uheating=1.36W(nominal)

ExperimentModel

Figure 6.6: (Experimental) Input Power to Output Torque small-signal response of the SMA

actuators, being uheating = a+ bsin(2πft), f = 2Hz. Source: The author.

power, Figure 6.6 compares the time-response of the model in Eq. (6.4) against experimental

data measured using the setup in Figure 6.2. By measuring the output torque generated by the

power signal of the form: uheating = a + bsin(2πft), it is possible to compare how the model

behaves for both nominal and operation modes of the SMAs. Nominal operation is when the

DC bias of input power is a = 1.36W whereas overloaded mode is when a = 3.06W . The small-

signal amplitude is defined by the parameter b which in both cases corresponds to b = 2mW .

The driven frequency of the results shown in Figure 6.6 corresponds to f = 2Hz. Similar results

have been observed with other ranges of driven frequencies.

6.3.3 Data summary

Table 6.1 summarizes the average numerical values of SMA actuation performance in terms of:

i) actuation speed, ii) input heating power, and iii) output torque.

The maximum allowed input current to drive SMA motion is set to Imax = 550 − 600mA.

This limitation is defined in order to avoid the SMAs to over-heat and therefore to over-stress.

In Table 6.1, simulation data indicates that the resultant nominal joint speed of 0.3deg/ms

results from applying an input heating current bias of I = 350mA (input power of 1.04W ).

During the experiments a slightly increase in power is required to maintain a similar joint

speed (0.26deg/ms). The difference in power consumption is due to non-modeled joint friction

and also from non-modeled membrane-load effects during wings extension. In spite of this, the

129

Page 162: BaTboT: a biologically inspired flapping and morphing bat robot

6.4 SMA Resistance-to-Motion relationship (RM)

Table 6.1: Summary of SMA actuation performance.

Parameter Theoretical1 Simulation2 Experimental3

Nominal

Joint speed q3 [deg/ms] 60/300 60/200 ∼ 60/225

Input heating power uheating [W ] 0.26 1.04 0.87 - 1.5

Output torque τ3 [Nm] 0.0012 0.0007 0.008

Overloaded

Joint speed q3[deg/ms] —- 60/100 ∼ 60/120

Input heating power uheating [W ] —- 2.57 2 - 3.06

Output torque τ3 [Nm] —- 0.0027 0.020

1 Nominal values provided by Migamotor’s model NanoMuscle RS-70-CE 1131 (96).2 Simulation results from Figure 4.22a (joint speed and input heating power), and Table 4.5 (output torque)3 Experimental results from Figures 6.5 (joint speed) and 6.6 (input power and output torque).

0 10 20 30 40 50 601

2

3

4

5

6

7

8

9

R [

Oh

ms]

θ3 [deg]

RM, To=23oC

RM, To=22.83oC

RM, To=22.75oC

RM, To=22.7oC

RM, To=22.6oC

R sma

Ω []

q3 deg[ ]

SMAs

Vair

Figure 6.7: (Experimental) Resistance-Motion (RM) linear relationship between SMA electri-

cal resistance change (Rsma) and the angular motion generated at the elbow joint (q3). Small

variations in ambient temperature (To) modify the RM relationship. The inset shows BaTboT

in the wind-tunnel. Ambient temperature has been measured with a MS 1000-CS-WC tempera-

ture sensor supplied by ATS (http://www.qats.com/Products). Changes in Rsma are constantly

measured during the experiment. Source: the author.

important issue concerns to keeping the input current below 600mA. Similar results have been

obtained during overloaded operation. Limitations in input power should be driven by a proper

SMA controller. Section 6.5 will present the control approach.

130

Page 163: BaTboT: a biologically inspired flapping and morphing bat robot

6.4 SMA Resistance-to-Motion relationship (RM)

6.4 SMA Resistance-to-Motion relationship (RM)

Although SMAs are mostly used for actuation, they also have sensing capabilities. Several

properties of the SMAs change as it undergoes the martensite phase transformation. Among

these properties is the resistivity that decreases as the temperature of the wire increases and

hence its phase transforms to austenite. In the proposed control scheme (see Figure 6.1), SMA

electrical resistance (Rsma) is the only property measured, and the inner morphing-controller

is, in effect, servo’ing electrical resistance to follow a commanded profile. One might expect

that resistance change is related to motion change, however, there is no direct measurement

of motion to evaluate the conditions under which this Resistance-Motion (RM) relationship

is valid. This fact suggests to analyze how the RM is affected when both ambient (To) and

SMA temperatures (T ) change. Even small variations in temperature modify the RM function’s

slope. The following section details on this issue.

RM sensitivity upon temperature change

At ambient temperature (To = 22.7oC) the Resistance-Motion function satisfies the linear

equation:

q3 = 10(8.5−Rsma) (6.5)

The function in Eq. (6.5) is determined by measuring the changes in SMA electrical re-

sistance during the heating phase of the wire. By also measuring the elbow joint rotation is

possible to determine a relationship between resistance and angular motion. This resistance-

angle relationship is linear because the martensite fraction is kinematically coupled to the

rotation, and the martensite fraction is what drives the resistance changes. Figure 6.7 shown

the experimental results. While the applied voltage and current to the SMA both change with

the wire strain in a hysteretic fashion, the resistance of the SMA wire (Rsma) changes almost

linearly with the angular motion q3. Several experiments at ambient temperature confirm the

reliability of the linear function.

However, the airflow generated by the flapping motion of the wings and also the airspeed

during flight might cause the SMA temperature to decrease, thus changing the RM function.

In Figure 6.7 it has been observed that variations in temperature lead to changes in resistance

decrease rate. Interestingly, changes were only produced in the slope of the curve, remained

the function linear. This phenomenon enables to state the RM relationship satisfy the model:

131

Page 164: BaTboT: a biologically inspired flapping and morphing bat robot

6.5 Morphing-wing control (inner loop)

q3 =M−1(Rsma − b), being M the slope of the curve. Using this model is possible to estimate

the angular motion q3 by measuring Rsma and calculating the parameters M and b.

6.5 Morphing-wing control (inner loop)

Figure 6.8 expands the inner control loop previously shown in Figure 6.1. It details how the

heating power signal uheating is generated such as mentioned in step 5 of Algorithm 3, cf. Eq.

(6.6). Two mechanisms: anti-slack and anti-overload complement the morphing-wing controller.

Two control strategies have been developed and implemented as morphing controllers: i) sliding-

mode control with estimated force reference, and ii) PID position control. In the following, both

controllers will be derived and compared.

AntagonisticSMA mechanism of the wings

Morphing controller-+ [ref]

[Uheating]SMA actuation

++ Ƭ3

Anti-slack mechanism

Anti-overload mechanism

++

[feedback]

u_high

u_low

u3

Figure 6.8: General scheme for the inner morphing-wing control. Source: the author.

uheating = u3 + ulow + uhigh,∈ �2, (6.6)

In Eq. (6.6) u3 refers to the morphing-wing control signal and ulow, uhigh are the lower and

upper values of input power that are regulated by the anti-slack and anti-overload mechanisms

respectively. Both mechanisms are used for both sliding and PID controllers.

6.5.1 Sliding-mode control

Figure 6.9 expands the inner loop shown in Figure 6.8. It completely details the components

required for the sliding-mode control method. The reference represents a torque profile to drive

elbow motion (τ3,ref ). Mapping from torque-to-motion requires the inertial model (FD). To

derive the control law equation u3 using the sliding-mode foundation, the following procedure

is applied:

132

Page 165: BaTboT: a biologically inspired flapping and morphing bat robot

6.5 Morphing-wing control (inner loop)

-+

Fsma_1

Fsma_2-1

SMA actuation

Resistancemeasurement

-+

Sliding mode controller

MIN

-+

0.03W

0.95

anti-slack mechanism

MAX

-+

3W

1.25

anti-overload mechanism

++

++

u_high

u_low

0.016/(0.35s+1)

0.016/(0.35s+1)

AntagonisticSMA mechanism of the

wings

u3

FD [q3,ref]

Inerial Model

q3 = M −1 Rsma − b( )RM function

q3

uheating τ 3τ difτ 3,ref HT( )−1 I3 q3,ref + ′Ksmce+ HT I3−1ξ3 +αsat s( )⎡⎣ ⎤⎦

Rsma =VsmaIsma−1

Psma = Isma2 Rsmainput power

Figure 6.9: Detailed inner loop of sliding-mode morphing-wing control with torque reference.

Source: the author.

1. Define the position error: e = q3,ref − q3 and its dynamics e.

2. Define the sliding surface S = e+K ′smce. S defines the dynamics that governs the system

behavior while sliding. K ′smc > 0 corresponds to the control parameters vector.

3. The sliding control signal u3 is designed by ensuring the references slide along the S

surface. A Lyapunov function is defined as: V = 0.5ST , S > 0.

4. The sliding control is chosen such as V = STS < 0, or STS ≤ −α |S| = −αST sgn(S),being sgn() the sign function, and α a positive scalar.

5. The sliding condition is S = −αsgn(s).

By differentiating S = e+K ′smce with respect to time:

S = q3,ref − q3 +K ′smce (6.7)

In Eq. (6.7) the term q3 governs the angular acceleration dynamics of the elbow joint. By

using the inertial model from Algorithm 3, this term corresponds to:

q3 = HT[I−13 (F3 − ξ3)

], (6.8)

where the 6x1 vector H allows for the projection of the spatial force of the elbow F3 onto

the axis of motion, ξ3 contains Coriolis and gyroscopic forces, and I3 is the six-dimensional

inertia matrix. By substituting Eq. (6.8) into Eq. (6.7):

S = q3,ref −HT[I−13 (F3 − ξ3)

]+K ′smce (6.9)

133

Page 166: BaTboT: a biologically inspired flapping and morphing bat robot

6.5 Morphing-wing control (inner loop)

By equaling Eq. (6.9) with the sliding condition S = −αsgn(s), and then isolating the force

term F3:

−αsgn(s) = q3,ref −HT[I−13 (F3 − ξ3)

]+K ′smce

F3 = (HT )−1I3[q3,ref +K ′smce+HT I−13 ξ3 + αsgn(s)]

(6.10)

By renaming u3 = F3, the sliding control law is derived from Eq. (6.10), as:

u3 = (HT )−1I3[q3,ref +K ′smce+HT I−13 ξ3 + αsgn(s)] (6.11)

Finally, to avoid chattering phenomenon, i.e., oscillations when the system approaches the

sliding region, the sign function is replaced by the saturation function sat(S), as:

sat(S) =

{sgn(S), S ≥ 0S/φ, |S| ≤ 0

}(6.12)

By adding Eq. (6.12) into (6.11), the sliding mode control law is:

u3 = (HT )−1I3[q3,ref +K ′smce+HT I−13 ξ3 + αsat(s)] (6.13)

The term ξ3 is defined in Eq. (4.10). The control gains K ′smc and α should be positive

scalars. Section 6.7 details the parameters values used for simulation and experimental testing

of the morphing-wing control using sliding-mode technique.

6.5.2 PID control

Figure 6.10 expands the inner loop shown in Figure 6.8 based on PID position control scheme.

The reference is the angular position profile of the elbow joint (q3,ref ). Using the linear model

in Eq. (6.4), the PID controller has been tuned using the Ziegler-Nichols methodology. The

PID transfer function is given by Eq. (6.14).

u3 = Kp

(1 +Kis

−1 +Kds), (6.14)

Where PID constants are: Kp = 35, Ki = 0.006, and Kd = 0.08. PID control is very simple

and most importantly, easy to program using the onboard limited processing capabilities of

BaTboT. It does not require complex computation and calculation. However, what really

makes the PID strategy the best choice for morphing-wing control driven by Shape Memory

Alloys recalls on mechanisms that enhance the PID control law. These mechanisms are defined:

i) anti-slack, and ii) anti-overload.

134

Page 167: BaTboT: a biologically inspired flapping and morphing bat robot

6.5 Morphing-wing control (inner loop)

-+

Fsma_1

Fsma_2-1

Resistancemeasurement

-+

PID

MIN

-+

0.03W

0.95

anti-slack mechanism

MAX

-+

3W

1.25

anti-overload mechanism

++

++

u_high

u_low

0.016/(0.35s+1)

0.016/(0.35s+1)

AntagonisticSMA mechanism of the

wings

35(1+0.006/s+0.08s)u3

[q3,ref]

q3

uheating τ 3τ dif

Rsma =VsmaIsma−1

Psma = Isma2 Rsmainput power

q3 = M −1 Rsma − b( )RM function

SMA actuation

Figure 6.10: Detailed inner loop of PID morphing-wing control with joint position reference.

Source: the author.

6.5.3 Mechanisms

Both sliding-mode and PID based inner control loops make use of two mechanisms that com-

plement the control law: anti-slack and anti-overload, cf. Figures 6.9 and 6.10 respectively.

These mechanisms allow:

• Anti-slack mechanism: it defines a minimum threshold of input heating power Pmin, that

ensures the inactive SMA wire does not cool completely. This lower value of input heating

power is ulow.

• Anti-overload mechanism: It is in charge of ensuring that the maximum input power does

not increase above an upper limit Pmax. This upper value of input heating power is uhigh.

The anti-slack mechanism

As explained in (67), the purpose of the anti-slack mechanism is to deal with the two-way

shape memory effect, which is mainly produced when the SMA wires extend upon cooling. The

passive SMA wire can develop a few millimeters of slack as it cools, which consequently affects

the accuracy of the control. The two-way shape memory effect becomes even more problematic

in the antagonistic arrangement of SMA actuators, and may lead to slower response and even

wire entanglement.

To avoid slack issues, the anti-slack mechanism defines a minimum threshold of input heating

power Pmin, that ensures the inactive wire does not cool completely. The improvement in

135

Page 168: BaTboT: a biologically inspired flapping and morphing bat robot

6.5 Morphing-wing control (inner loop)

actuation speed is due to the fact that the already-warm SMA wire can begin to pull as soon as

the heating current is raised, whereas a cold wire would first need to be raised to its operating

temperature. The mechanism compares the minimum value of the input power Psma of each

SMA with Pmin, ensuring that this applied power does not drop below the lower limit. In

the current system Pmin = 0.03W . After several simulations and experiments carried out

in section 6.7, it was empirically determined that by keeping 10% of the maximum applied

electrical current on the inactive SMA actuator, the mechanism works as expected. Thereby,

ulow = KsP , where Ks = 0.95 of the gain of the mechanism.

The anti-overload mechanism

The anti-overload mechanism is in charge of ensuring that the maximum input power does not

increase above an upper limit, defined as Pmax =∼ 2.57− 3W (calculated from the maximum

allowed input electrical current I =∼ 550 − 600mA, which was previously set in Figure 4.22,

Section 4.6.4). This approach avoids overheating the SMAs in case the controller delivers a

large amount of power. This input power saturation is due to uhigh = KoP , where Ko = 1.25

of the gain of the mechanism. It is important to highlight that the gains of each mechanism

(Ks, Ko) have been experimentally obtained to allow the elbow joint to rotate at a maximum

speed of ∼ 0.5deg/ms (cf. Figure 6.5).

6.5.4 Morphing-wing control algorithm

Algorithm 3 details how the inner-loop that drives the morphing-wing motion of the wings

operate. It runs typically at 30Hz, generating 12 samples per wing-stroke (12f), being f =

2.5Hz. In the first place, closing the inner loop requires the feedback of elbow motion q3.

Thanks to the RM function is possible to estimate q3 during the wingbeat cycle (cf. Figure

6.7). Having q3 estimated, the morphing-wing controller can operate (by using the sliding-

mode or PID strategies). Algorithm 3 works as follows: The outer -while- loop handles the

wingbeat cycle of the wings: downstroke and upstroke. During the downstroke, the SMA2

actuator is active whereas SMA1 is passive, allowing the wings to extend i.e., rotation of

elbow’s joint from ∼ 60o to 5o, (cf. Figure 6.11). Contrary during the upstroke, SMA1

turns active which allows the wings to contract. In step 1, an initial measurement of SMA

electrical resistance is measured and buffered in Rsma,i. This step is repeated at the beginning

of each wing stroke. The inner -while- loop handles the activation of each SMA actuator. This

loop ends when the -contracted - pin of the SMA actuator turns active, indicating it is fully

136

Page 169: BaTboT: a biologically inspired flapping and morphing bat robot

6.6 Attitude control (outer loop)

——————————————————-

ALGORITHM 3: morphing-wing control

——————————————————-

Initialize k ← 1, i ← 1, Mi ← 0.1, bi ← 8.5

while i < #wingbeat cycles do

1. Measure and buffer Rsma,i

2. Rsma,k ← Rsma,i

while CONTRACTED! = 1 do

3. Compute and feed-back q3,k ←M−1

i (Rsma,k − bi)

4. Calculate position error q3,ref − q3,k5. Run morphing control as uheating,k ← u3,k+

ulow,k + uhigh,k

6. k ← k + 1

7. Measure Rsma,k

end while

8. Buffer Rsma,f ← Rsma,k

9. Calculate new slope Mi+1 ← Rsma,f−Rsma,i

60o

10. Calculate new bi+1 ← Rsma,f −Mi60o

11. i ← i+ 1, k ← 1

end while

downstroke (SMA_2 active) upstroke (SMA_1 active)i=1, k=1

i++, k=1

k++

Measure

CONTRACTED =1

Measure

SMA_1

SMA_2

q 3,refdeg

[]

t s[ ]

Rsma,iRsma, f

Figure 6.11: Typical elbow joint reference

profile q3,ref during a wingbeat cycle (f =

1.25Hz). It details how Algorithm 3 works.

contracted. Inside this loop, at k = 1 (step 3), ebow joint is estimated by the RM model:

q3,k ← M−1i (Rsma,k − bi), where initial values are set to M = 0.1 and b = 8.5. Estimated

values of q3,k are feed-back to the control module aimed at calculating the position error of the

elbow joint: q3,ref−q3,k (step 4). The control signal uheating is generated by the morphing-wing

controller in step 5. uheating is a contribution of: morphing-wing control signal (u3), anti-slack

mechanism (ulow) and anti-overload mechanism (uhigh). Steps from 3 to 7 are repeated until

the end of each wingstroke, where new values of slopeM and term b are calculated based on the

final measurement of electrical resistance Rsma,f (steps 8 to 11). Simulations and experimental

results of the morphing-wing control can be found in section 6.7.

6.6 Attitude control (outer loop)

In bats, there is biological evidence that the inertial forces produced by the wings have a

significant contribution into the attitude movements of the animal, even more significant than

aerodynamic forces (24). In fact, bats perform complex aerial rotations by modulating solely

wing inertia (26). This means bats are able to change the orientation of the body during flight

without relying on aerodynamic forces and instead by changing the mass distribution of its

body and wings. Inertial forces are likely to be significant in bats because the mass of the wings

comprises a significant portion of total body mass, ranging from 11% to 33% and because wings

137

Page 170: BaTboT: a biologically inspired flapping and morphing bat robot

6.6 Attitude control (outer loop)

undergo large accelerations (49).

1

2

3

4

1

2

3

4

(a) (b)

3D joint points

Figure 6.12: a) In-vivo recordings of a bat land-

ing on the ceiling, cf. (26), b) closed-loop control

simulation of attitude maneuvering using the back-

steping+DAF strategy. Source: the author.

Thus, taking into account the effects

of wing inertia within the control law is

a key factor for the design of an attitude

controller of the proposed robot. The ap-

proach followed is aimed at defining proper

references that drive the modulation of the

wings’ shape in such a way to increment

inertial forces that generate propulsion. To

achieve this, an attitude control strategy

based on backstepping methodology plus

a function (DAF) to produce desired roll

and pitch angular accelerations are used.

The DAF function contains wing inertia

information that is provided by the iner-

tial model computation of the Equations of

Motion (EoM) of the robot. This enhanced

controller is called backstepping+DAF.

The novelty of the controller is: It pro-

poses a nonlinear control approach,

called backstepping+DAF, aimed at

improving the attitude response of the

MAV. Such enhancement is based on

the assumption (motivated by the cited biological studies in Chapter 3) that bats

efficiently generate forward thrust by means of inertia wing modulation, taking ad-

vantage of relevant wing-to-body mass ratio.. Figure 6.12 compares aggressive attitude

maneuvers executed by real bats (plot-a) and those obtained by the backstepping+DAF con-

troller from the simulator (plot-b). Joint trajectories of reference in plot-a has been extracted

from the specimen using high speed cameras (cf. (26)), and those references have been used by

the SimMechanics closed-loop simulator, cf. Annex 10.6, Figure 10.2. Annex 10.6 details the

SimMechanics closed-loop environment which is based on the open-loop environment previously

introduced in Figure 4.12. It complements the open-loop environment by adding both inner-

loop morphing-wing control and outer-loop attitude control. Because morphing-wing control

138

Page 171: BaTboT: a biologically inspired flapping and morphing bat robot

6.6 Attitude control (outer loop)

BacksteppingMotor

outer loop: attitude control R/C (flapping)

IMU

BaTboT model

+-

R/C (attitude)forward/turning

DAF

inner loop: morphing control

[q1…q6]R,L

wingtrajectories

[roll, pitch]

q1,refφ d ,θ d⎡⎣ ⎤⎦φ,θ[ ]

uφ ,uθ⎡⎣ ⎤⎦

wristmapping

φ,φ,θ ,θ⎡⎣ ⎤⎦

uθ = q2,ref

uφ = q3,ref

-+

Fsma_1

Fsma_2-1

Resistancemeasurement

-+

PID

MIN

-+

0.03W

0.95

anti-slack mechanism

MAX

-+

3W

1.25

anti-overload mechanism

++

++

u_high

u_low

0.016/(0.35s+1)

0.016/(0.35s+1)35(1+0.006/s+0.08s)

u3

q3

uheating τ dif

Rsma =VsmaIsma−1

Psma = Isma2 Rsmainput power

q3 = M −1 Rsma − b( )RM function

SMA actuation

τ1τ 2

τ 3τ 4τ 5τ 6

Figure 6.13: Detailed outer and inner loops: complete description of the Flight Control Archi-

tecture (FCA). Source: the author.

can be based on sliding-mode technique or PID technique, the former is fully described in Fig-

ure 10.2 whereas de latter in Figure 10.2. The SimMechanics closed-loop environment

is not only used for simulation purposes but also for experimental testing. It sends

and receives commands to/from the Arduino-nano board that is placed on the robot. Details

on software-to-hardware control interplay can be found in Figure 6.14. In addition, Figure

6.13 expands the outer-loop attitude control from Figure 6.1 and shows the Flight Control

Architecture (FCA) is more detail.

6.6.1 Backstepping+DAF

DAF: Desired Angular acceleration Function (φd, θd)

To formulate the DAF terms (φd, θd), the attitude data (θ, θ, φ, φ) that is feedback by the IMU

and the wing joint trajectories of reference (q) are required, as shown in Figure 6.13. Therefore

DAF terms can be written as a function of attitude and wing modulation data as:

φd = f(φ, φ, q, q)

θd = f(θ, θ, q, q)(6.15)

The desired roll and pitch angular accelerations (φ, θ) are components of the six-dimensional

body accelerations Vb that are produced by the inertial forces FT . The definition of Vb with

139

Page 172: BaTboT: a biologically inspired flapping and morphing bat robot

6.6 Attitude control (outer loop)

respect to the body frame {b} requires the computation of the inertial model from Algorithm

1, cf. step 5, and yields:

Vb =

(Ib +

[R0,b

0∑i=6

(Ri+1,iPTi,i+1Ii,R)+R0,b

0∑i=6

(Ri+1,iPTi,i+1Ii,L)

])−1

FT (6.16)

In Eq. (6.16), Ib is the spatial inertia of the robot’s body calculated with respect to the

body frame {b}, whereas the term0∑i=6

Ii expresses the propagation of wing inertias of both

wings onto the base frame {0}. Subscripts R and L refer to the right and left wing respectively.

The term R0,b applies a rotation to express wing inertias with respect to the body frame {b}.Finally, DAF terms are defined and expressed with respect to the body frame {b} as:

φd =[1 0 0 0 0 0

]Vb,

θd =[0 1 0 0 0 0

]Vb

(6.17)

Backstepping+DAF control

This section shows roll control derivation for uφ. Pitch control derivation (uθ) follows the same

procedure. The first step is to define the roll tracking error e1 and its dynamics (derivative

with respect to time):

e1 = φd − φ

e1 = φd − ωy(6.18)

The term φd corresponds to the desired roll trajectory profile, whereas φ is the roll angle

measured by the IMU sensor. A positive definite Lyapunov function (L) is used for stabilizing

the tracking error e1, as:

L (e1) =e212 , (6.19)

To regulate the behavior of the angular velocity ωy from e1 in Eq. (6.18), a second tracking

error e2 is defined, as:

e2 = ωdy − ωy (6.20)

The desired behavior for e2 is defined as:

ωdy = c1e1 + φd + λ1∫e1 , (6.21)

140

Page 173: BaTboT: a biologically inspired flapping and morphing bat robot

6.6 Attitude control (outer loop)

where c1, λ1 are positive constants and∫e1 is the integral of the roll tracking error. In

other words, ωdy is considered as a virtual control law that governs the behavior of e2. Now,

substituting ωdy into Eq. (6.20) and differentiating e2 with respect to time (note that ωy = φ):

e2 = c1e1 + φd + λ1e1 − φ (6.22)

From Eq. (6.20) ωy = ωdy − e2. Substituting ωy into e1 from Eq. (6.18) gives:

e1 = φd − ωdy + e2 (6.23)

Now, substituting ωdy from Eq. (6.21) into (6.23):

e1 = φd − (c1e1 + φd + λ1∫e1) + e2

e1 = e2 − c1e1 − λ1∫e1

(6.24)

Finally, substituting e1 from Eq. (6.24) into Eq. (6.22):

e2 = c1(e2 − c1e1 − λ1∫e1) + φd + λ1e1 − φ (6.25)

By expressing φ in terms of the pitching torque τφ and the moment of inertia of the robot’s

body Ixx about the x axis of the body frame {b} φ = τφ/Ixx and defining uφ = τφ, Eq. (6.25)

can be written as:

uφ = Ixx[c1(e2 − c1e1 − λ1∫e1) + φd + λ1e1 − e2] (6.26)

Replacing e2 = −e1 −λ2e2 into Eq. (6.26), the backstepping+DAF control law in charge of

roll regulation is:

uφ = Ixx[e1(λ1 − c21 + 1) + e2(c1 + λ2)− c1λ1∫e1 + φd] (6.27)

The parameters of the controller, λ1, c1, λ2 > 0 are defined in Table 6.2. By applying the

same procedure, the backstepping+DAF control law in charge of pitch regulation is given by:

uθ = Iyy[e3(λ3 − c22 + 1) + e4(c2 + λ4)− c2λ3∫e3 + θd] (6.28)

By substituting the DAF terms φd, θd into Eq. (6.27) and (6.28):

uφ = Ixx[e1(λ1 − c21 + 1) + e2(c1 + λ2)− c1λ1∫e1 +

[1 0 0 0 0 0

]Vb

uθ = Iyy[e3(λ3 − c22 + 1) + e4(c2 + λ4)− c2λ3∫e3 +

[0 1 0 0 0 0

]Vb

(6.29)

141

Page 174: BaTboT: a biologically inspired flapping and morphing bat robot

6.6 Attitude control (outer loop)

Stability Analysis

The following candidate Lyapunov function has been chosen:

L = 0.5(e21 + e22 +∫e1

2) (6.30)

Differentiating Eq. (6.30) with respect to time, having e1 = e2 − c1e1 − λ1∫e1 and e2 =

−e1 − λ2e2 gives:

L = e1e1 + e2e2 + e21 ≤ 0= e1(e2 − c1e1 − λ1

∫e1) + e2(−e1 − λ2e2) + e21

= −c1e21 − λ2e22 ≤ 0

(6.31)

The fact that Eq. (6.31) fulfils L ≤ 0, ∀(e1, e2) ensures the boundedness of e1, its integral∫e1, and e2. Hence, the reference angular value φd and the roll angle φ are also bounded

due to e1 = φd − φ. The boundedness of e1 implies that the virtual law ωdy is bounded as

well, which consequently makes the error dynamics e2 and the DAF term φd also bounded.

Furthermore, global asymptotic stability is also ensured due to the positive definition of L, in

which L(e1, e2) < 0, ∀(e1, e2) = 0, and L(0) = 0 (by applying LaSalle’s theorem).

6.6.2 Attitude control algorithm

The attitude control procedure based on the proposed backstepping+DAF methodology is

detailed in Algorithm 4. The algorithm is based on the Flight Control Architecture (FCA)

detailed in Figure 6.13.

ALGORITHM 4: Backstepping+DAF attitude control

——————————————————-

1. Read filtered data from IMU: φ, φ, θ, θ

2. Compute Vb from Algorithm 1 using IMU readings.

3. Calculate DAF terms: φd, θd

4. Run backstepping+DAF control: uφ, uθ

5. Run inner-loop morphing-wing control Algorithm 3.

The computation of the inertial model from Algorithm 1 is required for calculating the DAF

terms that compose the attitude controller. The inputs to the inertial model are the wing joint

trajectories of reference (q, q, q) and the filtered attitude IMU measurements (φ, φ, θ, θ). The

output of the inertial model is the spatial body accelerations Vb that are produced by wing

forces that act on the center of mass of the robot. DAF terms are then extracted from Vb in the

form of desired angular acceleration commands which drive roll and pitch behaviour (φd, θd).

142

Page 175: BaTboT: a biologically inspired flapping and morphing bat robot

6.7 Simulation and experimental results

Both DAF terms are added into the backstepping control law aimed at improving roll and pitch

regulation (uφ, uθ). Improvements are expected in terms of increments of inertial thrust due

to more efficient modulation of the wings that are driven by the backstepping+DAF control

outputs. The following section presents preliminary results of the Flight Control Architecture

(FCA) on BaTboT.

6.7 Simulation and experimental results

Simulations are carried out using the closed-loop Simmechanics Matlab environment shown

in Annex 10.6, whereas experiments are conducted by using the closed-loop setup described

in Figure 6.14. The setup shows the software and hardware interplay of the Flight Control

Architecture (FCA) modules from Figure 6.13.

16-bit ADC + anti-aliasing filter9620-05-ATI

wind-tunnel testbed withforce sensor load

SMA driver

arduino-nano SMA resistancemeasurement

FT

user

Matlabenvironment

[u1]

[u3,R u3,L]

(morphing)

[uheating,L,R]

[u1]

(flapping)

[Rsma,L,R]

[roll pitch ]φ,φ,φ,θ ,θ ,θ

[R_feedback]

[Isma]

IMUφ θ

(force)

Figure 6.14: Experimental setup using the wind-tunnel of Brown University: Flight Control

Architecture (FCA). Source: the author.

BaTboT is equipped with onboard processor based on Arduino technology which uses a

PIC18F2680 that mainly receives the commands from the external PC (Matlab environment)

via serial connection. These commands are the control outputs for: flapping u1 and morphing-

wing modulation u3,R,L (subscripts R and L refer to the right and left wing respectively).

The SMA driver is based on a MOSFET transistor that receives the control command u3

and generates the driving electrical current signal to operate each SMA actuator of the wings.

143

Page 176: BaTboT: a biologically inspired flapping and morphing bat robot

6.7 Simulation and experimental results

BaTboT is also equipped with onboard IMU that feedbacks attitude measurements: φ, φ, θ, θ.

These signals are filtered inside the Arduino processor using a Kalman filter before feedback to

the PC. Also, SMA electrical resistance changes are also measured (Rsma) and feedback to the

PC. Finally, a cell load within the wind-tunnel equipped with a force sensor allows for getting

measurements of 6D forces produced at the center of mass of the robot. A 16-bit DAC with

embedded 1st order anti-aliasing filter is used for that purpose.

Attitude outer loop runs at 20Hz based on IMU readings, whereas morphing-wing inner loop

runs at 30Hz based on SMA resistance change readings. Details on these loops are depicted in

Figure 6.13.

6.7.1 Variables and parameters

The list of parameters used for both simulation and experimental testing are consigned in Table

6.2. It details morphological, modeling and control parameters.

Table 6.2: List of robot’s parameters: morphological, modeling, control.

Parameter (unit) Value

Body mass Mb (g) 125

Extended wing length B (m) 0.245

Body width lm (m) 0.04

Body inertia tensor diagonal [Ixx, Iyy , Izz ] (gcm2) [1, 0.07, 0]

Extended wing spanS = lm + 2B (m) 0.53

Extended wing area Ab (m2) 0.05

Humerus lengthlh (m) 0.055

Humerus inertia tensor diagonal [Ixx, Iyy , Izz ] [gcm2] [0.03, 0.37, 0.93]

Humerus position vector to CM s2,cm (m) [0.0275, 0, 0]

Radius lengthlr (m) 0.070

Radius inertia tensor diagonal [Ixx, Iyy , Izz ] (gcm2) [0.07, 0.92, 0.37]

Radius position vector to CMs3,cm (m) [0.035, 0, 0]

Air density ρair (Kg/m3) 1.2

Lift coefficient CL at angle of attack AoA = 9o 1.5

drag coefficientCDat angle of attack AoA = 9o 0.152

Average lift force FL at angle of attack AoA = 9o (N) 0.97

Average drag force FD at angle of attack AoA = 9o (N) 0.099

Sliding mode morphing-wing controller parameters K′smc, α 2.45, 1.1

PID morphing-wing controller parameters Kp, Ki, Kd 35, 0.006, 0.08

Anti-slack mechanism gain Ks 0.95

Anti-overload mechanism gain Ko 1.25

Backstepping+DAF attitude control (roll) [λ1, c1, λ2] [1.87, 2.1, 0.02]

Backstepping+DAF attitude control (pitch)[λ3, c2, λ4] [4.5, 2.5, 0.02]

144

Page 177: BaTboT: a biologically inspired flapping and morphing bat robot

6.7 Simulation and experimental results

6.7.2 Morphing-wing control response

In this section both derived sliding-mode and PID methods are compared aimed at determining

the best performance for driving SMA actuators. As introduced at the beginning of this chapter

the control goal relies on achieving the fastest wing modulation via SMA actuation

without compromising power consumption and yet achieving a feasible SMA output

torque. Both control strategies are evaluated in that regard.

Sliding-mode response

The control law derived in Eq. (6.13) whose diagram block is detailed in Figure 6.9 is first tested

on simulation. Figure 6.15 shows the results. Sliding-mode control was initially proposed

under the thought of being capable of driving faster SMA contraction thanks to the energy

management ensured by the control law. With that in mind, a sinusoidal joint position reference

for elbow rotation (q3,ref ) was set to a frequency of 4Hz. Plot-a in Figure 6.9 shows the tracking

results. A first problem is observed in terms of error tracking which causes a delay in tracking

the reference. Several simulation results have shown similar response. The problem is caused

by the saturation function of the control law (cf. Eq. 6.12). This function avoids chattering

phenomenon, which means oscillations when the system approaches the sliding region of the

surface S, and shown in plot-b. This technique does not ensure the asymptotic stability of the

system and compromises an accurate response by smoothness. As a consequence plot-a shows

a smooth tracking but unfortunately a large tracking error.

In terms of power consumption the sliding-mode does well in simulation. Plot-c in Figure

6.15 details the driven current consumed by each SMA actuator. Note how the anti-slack

mechanism ensures the input heating power does not drop below the lower threshold of 0.03W ,

which in terms of electrical current corresponds to 0.06A. This approach allows to keep the

SMA wires away from completely cooling which in turns causes that the wires can respond

faster under activation/contraction. Using the sliding-mode technique a maximum current of

0.2A is required to actuate the SMA at the driven frequency. This is a good result taking

into account each SMA contracts in 100ms. In Figure 6.16 experiments are conduct aimed at

verifying the simulation results.

As mentioned in Section 6.1, one of the control goals is to evaluate efficiency as a tradeoff

between SMA output torque (τ3) and input power (Psma = I2smaRsma). By using the inertial

model to map torque references (τ3,ref ) to joint position references (q3,ref ) and by measuring

145

Page 178: BaTboT: a biologically inspired flapping and morphing bat robot

6.7 Simulation and experimental results

(a)

q3 [rad] -1 -0.5 0 0.5 1 1.5

-2

q3 [r

ad/s

]

.

-1

0

123

4

5

6

S=0

t [s]

S

e

e.

sliding surfacesat(S) function

0t [s]

0.2 0.4 0.6 0.8 1

Isma [

A]

0

0.05

0.1

0

0.15

0.2

0

Isma_1Isma_2

(b)

q3

q3,ref

0 0.1 0.2 0.3 0.4 0.5 0.6t [s]

-1

-0.5

0

0.5

1

(c)

q3 [r

ad]

Anti-slack mechanism(Imin=0.06A)

Figure 6.15: (Simulation) morphing-wing based on sliding-mode response: a) control tracking

given a sinusoidal joint trajectory (q3,ref ) at wingbeat frequency of f = 4Hz, b) phase plane of

the sliding surface upon sinusoidal input from plot (a), c) electrical current (Isma) to drive each

SMA actuator in the antagonistic configuration. Source: the author.

the output torque (τ3) using the setup shown in Figure 6.2, experimental results regarding

torque tracking are depicted in Figure 6.16a. It turns out the sliding-mode technique causes

fatigue issues on the SMA wires due to large peaks of current delivered. This can be clearly

observed in plot-a. Note how during the second wingbeat (t > 0.6s) the output torque τ3

saturates even under the presence of electrical heating. This causes large errors in tracking that

can be appreciated in Figure 6.16b. The saturation of output torque occurs due to a physical

decrease of performance of the SMAs which in turns limits the angle range of motion of the

elbow joint q3. At the beginning of the experiment it was thought that these large errors in

joint tracking were caused by the saturation function of the sliding-mode control, similarly with

those observed from the simulations, cf. Figure 6.15a. Nonetheless, after removing the sat(S)

function from the control law, the tracking error did not improved. Figure 6.16b compares the

joint position tracking with and without the saturation function. Therefore, it was discovered

that fatigue issues on the SMAs caused by the controller was the cause of these errors. Plot-b

contains the joint tracking curve of a wingstroke corresponding to the interval between 0.6s and

1s, which fatigue issues are clearly observed from the plot-a. As mentioned fatigue is caused by

146

Page 179: BaTboT: a biologically inspired flapping and morphing bat robot

6.7 Simulation and experimental results

Isma_1Isma_2

Isma [

A]

t [s]

τ3 [N

m]

τ3,ref

τ3

(a)

q3 [d

eg]

0 0.2 0.4 0.6 0.8 1 0.4 0.45 0.5 0.55 0.6 0.65 0.7 0 00.75 0.8t [s]

(b)

0

0.02

-0.02

0.01

-0.01

t [s]

0

12

24

36

48

60q3 with sat(S)q3 without sat(S)

q3,ref

(c)

Fatigue

0

0.25

0.5

0.75

1

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

τ3,ref

τ3

Figure 6.16: (Experimental) morphing-wing based on sliding-mode response: a) SMA output

torque τ3 given a force reference τ3,ref , b) control tracking given a bio-inspired joint trajectory

q3,ref at wingbeat frequency of f = 2.5Hz; it shows comparison of sliding-mode response with

and without the saturation function sat(S) within the control law, c) electrical current (Isma) to

drive each SMA actuator in the antagonistic configuration Source: the author.

large applied electrical currents delivered by the controller. Plot-c describes the current profile

applied to each SMA actuator of the antagonistic arrangement. It corresponds to the first

wingstroke from 0s to 0.4s. Peaks up to 1A causes the SMA to fatigue. This causes an instant

decrease of output torque (cf. plot-a) and therefore a decrease in motion range. The decrease in

motion range is clearly observed in plot-b, causing the tracking error mentioned before. Fatigue

might cause a fatal damage of the shape memory effect. Further simulations and experiments

regarding sliding-mode control can be found in (97), whereas PID control response is presented

in the following.

147

Page 180: BaTboT: a biologically inspired flapping and morphing bat robot

6.7 Simulation and experimental results

0 0.5 1 1.5 20

20

40

60

t [s]

q 3[deg]

0 0.5 1 1.5 20

0.2

0.4

0.6

t [s]

I sma[A

]

0 0.5 1 1.5 20

20

40

60

t [s]

q 3[deg]

0 0.5 1 1.5 20

0.2

0.4

0.6

t [s]I sm

a[A]

0 0.5 1 1.5 20

20

40

60

t [s]

q 3[deg]

0 0.5 1 1.5 20

0.2

0.4

0.6

t [s]

I sma[A

]

q3,refq3

q3,refq3

q3,refq3

Isma,1Isma,2

Isma,1Isma,2

Isma,1Isma,2

Anti-slack mechanism(Imin=10%*Imax)

Anti-overload mechanism

(a) (b)

Figure 6.17: (Simulation) morphing-wing based on PID response: a) PID control tracking given

a sinusoidal profile q3,ref (above), square profile (medium), sawtooth profile (below), at wingbeat

frequency of f = 2.5Hz, b) electrical current (Isma) to drive each SMA actuator in the antagonistic

configuration. Source: the author.

PID response

PID control architecture was previously shown in Figure 6.10. The control law u3 in Eq. (6.14)

was tuned based on experimental identification of the plant, which in this case corresponds to

the SMA actuators. The identified plant were represented by a LaPlace first order function

shown inn Eq. (6.4).

The main advantage of the PID control relies on the less amount of power that delivers

to the SMAs compared to the sliding-mode. Firstly simulations are carried out to analyse

the tradeoff between joint tracking accuracy and delivered power. Figure 6.17a details the

simulation results taking into account three different profiles for describing the joint reference

q3,ref : sinusoidal, square and sawtooth. Figure 6.17b shows the corresponding input heating

current delivered by the PID controller. For instance it can be observed that the maximum

required current to drive both SMA actuators is always below 550mA, about the half than the

experiments carried out with the sliding-mode technique (cf. Figure 6.16c). In terms of tracking

the PID does well, specially for sinusoidal input references (cf. Figure 6.17a, above). Comparing

against the sinusoidal tracking achieved with the sliding-mode in Figure 6.15a, which phasing

148

Page 181: BaTboT: a biologically inspired flapping and morphing bat robot

6.7 Simulation and experimental results

0 2 4 6 8

0.005

0.01

0.015

0.02

0.025

t [s]

3 [Nm]

0 2 4 6 80

0.15

0.3

0.45

0.6

0.75

t [s]

I sma [A

]

0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.40

12

24

36

48

60

72

t [s]q 3 [d

eg]

q3,refq3

Anti-slack mechanism(Imin=10%*Imax) 0

Anti-overload mechanism(Imax=0.6A)

downstroke upstroke

0

(a)

(c)

(b)

f=2.5Hz

Figure 6.18: (Experimental) morphing-wing based on PID response: a) measured SMA output

torque τ3, b) PID control tracking given a bio-inspired joint trajectory q3,ref , c) electrical current

(Isma) to drive each SMA actuator in the antagonistic configuration. Source: the author.

problems are presented, the PID is able to accurate track the reference with a position error of

1.5%. In addition, there has not been observed any issues related to fatigue problems. However,

further experiments carried out in Chapter 7 section 7.2.1.2 will assess and discuss this issue in

more detail.

Experimental response of PID morphing-wing control is detailed in Figure 6.18. Note in

plot-a the measured output torque is maintained thanks to the proper regulation of the driven

current. Plot-c depicts the current profile in where is clearly appreciated the effects of both

anti-slack and anti-overload mechanisms. The anti-slack monitors the driven current does not

drop below 10% the maximum threshold aimed at speeding up the actuation frequency. In

this experiment the frequency is about 2.5Hz. On the other hand the anti-overload mechanism

monitors the driven current does not increase above the maximum threshold aimed at avoiding

fatigue issues. This is another advantage of PID control. Note how simulations reported

in Figure 6.17b (above) and experiments in Figure 6.18c show similar applied current limits

(below 600mA) that enable the proper wing modulation of joint q3. Contrary to the PID

control, the sliding-mode cannot take advantage of the anti-overload mechanism to ensure the

current limitation because it saturates the control output avoiding proper control tracking.

149

Page 182: BaTboT: a biologically inspired flapping and morphing bat robot

6.7 Simulation and experimental results

Without the anti-overload mechanism fatigue issues may occur, as shown in cf. Figure 6.16a.

Finally Figure 6.18b details the PID control tracking of the elbow joint reference q3. PID is able

to track the reference with neglectable error. Tracking errors are presented because the motion

of the elbow that is feedback to the controller has not been directly measured, instead and

estimation of the motion based on the change of SMA electrical resistance is the only variable

measured (cf. Algorithm 3: RM function). This estimation contains accumulative errors that

are presented during the elbow tracking, nonetheless these errors are not significant to decrease

the performance of the PID controller.

6.7.3 Attitude control response

This section only presents preliminary results of backstepping+DAF control in terms attitude

tracking. Attitude control results regarding the goal of incrementing net body forces thanks to

the wing modulation driven by the backstepping+DAF control law will be covered and discussed

in the main experimental results Chapter 7 section 7.4.

Backstepping has been widely applied to robust flight control problems (98). The key idea

of the backstepping design is to benefit from the desired dynamic state feedback that composes

the control law u = f(z, zd) (99). In this case, z corresponds to (φ) or (θ) respectively. The

backstepping does not only ensure global asymptotically stabilisation of φ and θ, it also allows

for the definition of a Desired angular Acceleration Function (DAF) to define the terms (φd)

and (θd). Basically two terms within the backstepping+DAF are key for attitude control: i)

integral action, and ii) DAF terms. Both can be observed from Eq. (6.29), where integral

action is represented by the term c1λ1∫e1 and DAF terms are φd, θd. The attitude error e1 is

expressed by e1 = φref − φ in the case of roll regulation (similarly with pitch). Integral action

is key for stabilisation purposes, whereas DAF terms are key for tracking purposes. Both are

combined into a single control law that enables BaTboT with accurate attitude performance.

Figures 6.22 and 6.20 show simulation and experimental respond of backstepping+DAF

control for stabilisation purposes. In Figure 6.22, both roll and pitch angles must be stabilise

to zero. Note roll angle (φ) has been set to 20o whereas pitch angle (θ) to −20o. The controller

is able to regulate the roll angle with a maximum control effort about 0.03Nm and the pitch

angle with a control effort about 0.15Nm. Comparing both values against the experimental

measurements carried out for the quantification of roll and pitch maneuvering from Figure 4.19,

one can note the similarities between simulation and experimental data. Refer to Table 4.7 for

numerical values at f = 1.3Hz. Experimental values indicate 0.04Nm for roll and 0.11Nm for

150

Page 183: BaTboT: a biologically inspired flapping and morphing bat robot

6.7 Simulation and experimental results

0 2 4 6 8 10−20

−15

−10

−5

0

5

t [s]

θ[d

eg]

0 2 4 6 8 10−0.04

−0.03

−0.02

−0.01

0

0.01

t [s]

uφ [

Nm

]

0 2 4 6 8 10−0.05

0

0.05

0.1

0.15

0.2

t [s]

uθ [

Nm

]

0 2 4 6 8 10−5

0

5

10

15

20

t [s]

φ [

deg

]

φref

φ

θref

θ

Figure 6.19: (Simulation) backstepping+DAF attitude stabilization. Roll (φref ) and pitch (θref )

references are set to zero while initial attitude position is set to 20o and −20o for roll and pitch

respectively. Control efforts are also shown. Source: the author.

0 2 4 6 8−2.5

−2

−1.5

−1

−0.5

0

0.5

1

1.5

t [s]

φ [

deg

]

0 2 4 6 8−3

−2

−1

0

1

2

3

4

t [s]

θ [

deg

]

θref

θ

φref

φ

Figure 6.20: (Experimental) backstepping+DAF attitude stabilization. Roll (φref ) and pitch

(θref ) references are set to zero. The wind-tunnel airspeed has been set to 2ms−1 and the controller

must keep both angles close to zero. Source: the author.

pitch. In Figure 6.20, both roll and pitch angles must be stabilise to zero in spite of disturbances

produced by the wind-tunnel airspeed, which it has been set to 2ms−1. The oscillations around

the set points are caused by the natural wingbeat motion (f = 1.5Hz) and the aerodynamic

loads. Position errors less than 3o are observed from the experiment.

In terms of tracking, backstepping+DAF also does well. Figure 6.21 show simulation results

151

Page 184: BaTboT: a biologically inspired flapping and morphing bat robot

6.7 Simulation and experimental results

0 0.5 1 1.5 2 2.5 3−2

−1.5

−1

−0.5

0

0.5

1

1.5

2

t [s]

φ [

deg

]

0 0.5 1 1.5 2 2.5 3−6

−4

−2

0

2

4

6

t [s]θ

[deg

]

φref

φ

θref

θ

Figure 6.21: (Simulation) backstepping+DAF attitude tracking. Roll (φref ) and pitch (θref )

sinusoidal references are tracked. Source: the author.

0 1 2 3−2

−1

0

1

2

φ [

deg

]

t[s]

0 1 2 3−10

−5

0

5

θ [

deg

]

t[s]

0 1 2 30

0.01

0.02

0.03

erro

r

t[s]

φref

at 0ms−1

φ at 0ms−1

φref

at 5ms−1

φ at 5ms−1

θref

at 0ms−1

θ at 0ms−1

θref

at 5ms−1

θ at 5ms−1

φ error at 5ms−1

φ error at 0ms−1

0 5 10 150

0.005

0.01

0.015

0.02

erro

r

t[s]

θ error at 5ms−1

θ error at 0ms−1

1 2 3

Figure 6.22: (Experimental) backstepping+DAF attitude tracking. Roll (φ) and pitch (θ) pro-

files are tracked for wind-tunnel airspeeds of 0 and 5ms−1. Tracking errors are measured. Source:

the author.

regarding roll (φref ) and pitch (θref ) sinusoidal references. Tracking errors are quantified during

the experiments shown in Figure 6.22, which also depicts attitude tracking for different wind-

tunnel airspeeds profiles; 0 and 5ms−1 respectively. Tracking errors less than 3% are observed.

152

Page 185: BaTboT: a biologically inspired flapping and morphing bat robot

6.8 Remarks

6.8 Remarks

Regarding morphing-wing control, PID has demonstrated to be sufficient for achieving accurate

SMA actuation. It enables the SMA actuators to actuate at a maximum frequency of 2.5Hz

while ensuring accurate tracking and proper power consumption. Similar conclusions can be

found in (9). In fact, before (9) and (67), PID controllers were considered slow and inefficient,

and investigations regarding PID with high gains were never carried out. In this thesis, we have

also demonstrated that the power-to-force AC response of a NiTi SMA wire can be accurate

modeled as a first order low-pass filter, which transfer function was experimentally identified

by carrying out a frequency response analysis over the range of interest. This model is shown

in Eq. (6.4). Because the AC response of the SMA muscles resemble a first order low-pass

filter, the use of a PID seems to be an adequate solution. It is easy to program it onboard the

robot’s microprocessor and it can also be easily tuned under both simulation and experiments.

Basically, the main advantage of the PID control relies on the less amount of power that delivers

to the SMAs thanks to the incorporation of both mechanisms that monitors upper and lower

thresholds of SMA input power.

Regarding attitude control, the proposed backstepping+DAF controller has demonstrated

to be key for achieving proper response in terms of attitude stabilisation and tracking. Integral

action cancel steady errors whereas DAF terms improve on tracking response. However, the

key role of the DAF terms is to include wing inertia information that enables proper wing mod-

ulation. This issue has not been evaluated in this section but it will be completely introduced

in the following Chapter 7.

153

Page 186: BaTboT: a biologically inspired flapping and morphing bat robot

7

General experimental results

The fastest maneuvers in flying animals could be reproduced

in man made flying vehicles.

7.1 Overview

This chapter presents further experiments aimed at discussing the methods (modelling and con-

trol) proposed in this thesis and assessing their potential for developing bio-inspired morphing-

wings bat-like aerial robots. Along the previous chapters, it is has been presented a detailed

workflow that describe all the processes involved in that goal, from the analysis of biological

data that inspire and define the design of BaTboT, to the kinematics, dynamics, aerodynamics,

actuation and control methods that enable BaTboT to behave like its biological counterpart.

7.1.1 Methods and goals

Experiments are categorised in three areas:

1. Control performance. It evaluates both morphing-wing and attitude control response.

The goal of the experiments is twofold: (i) to assess the performance of SMA actuators

in terms of accuracy, limitations and impact into the proper modulation of the morphing-

wings. (ii) to evaluate how the proposed attitude controller enables accurate forward and

turning flight.

2. Aerodynamics. The goal is to demonstrate the impact of both morphing-wing and attitude

controllers into the proper generation of aerodynamic forces. It also discusses how to

154

Page 187: BaTboT: a biologically inspired flapping and morphing bat robot

7.1 Overview

robotic-arm(angle of attack)

high-speed camera

6D force sensor

high-speed camera

wind-tunnel mount-platform

(a)

(b)

Wind-tunnel test section

Figure 7.1: a) setup for morphing-wing testing. b) wind-tunnel setup for dynamics, aerodynamics

and control testing. Source: the author.

induce accurate aerodynamic behaviour for similar robots, not necessarily bat-like, based

on kinematics, dynamics and control parameters.

3. Efficient Flight. Efficiency is evaluated in terms of net force production. The goal is to

quantify how net forces can be increased by modulating wing inertia in a proper way.

Here, the hypothesis introduced at the beginning of this thesis is demonstrated, which

states that by including wing inertia information into the attitude controller, efficient

wing modulation that induces proper attitude behaviour is achieved.

155

Page 188: BaTboT: a biologically inspired flapping and morphing bat robot

7.2 Control performance

7.1.2 The wind-tunnel setup

Quantification of dynamics and aerodynamics data requires a complex setup. For this, the

Brown University’s wind-tunnel facility is used. Along the chapters of this thesis, a brief

description of the wind-tunnel components have been presented, as shown in Figure 6.14, how-

ever, a detailed description of the entire setup has not been shown. To this purpose, Figure

7.1 describes the main experimental setups used in this thesis. Firstly, the design of a highly

articulated morphing-wing required the use of a setup specially conceived for assessing SMA ac-

tuation. This setup is depicted in Figure 7.1a and enables half-part of the robot to be mounted

in (right wing). Secondly, Figure 7.1b shows the wind-tunnel setup. It enables the entire robot

to me mounted on a 6D force sensor that measures dynamics and aerodynamic forces. Wind

tunnel experiments have been conducted at the Brown University Breuer’s lab facility, which

is a closed-loop circuit with a test section measuring 3.8m in length and a cross-section of 0.60

by 0.82m (height X width). The wind-tunnel maximum airspeed is 20ms−1 and the stream

turbulence level of the tunnel is quite low (0.29% at 2.81ms−1). High-resolution CMOS cam-

eras (Photron 1024 PCI, resolution 1024x1024 pixels, lens 85mm, f/1.4) allow to capture all

markers useful for the kinematics extraction.

7.2 Control performance

This section complements experimental tests carried out in Chapter 6. The list of parameters

used for the following experiments remains the same than those introduced in Table 6.2. Also,

PID morphing-wing inner control loop and backstepping+DAF attitude outer control loop

structures remain the same than those introduced in Figure 6.13.

7.2.1 Morphing-wing control

To assess the performance of the Morphing-wing control, measurements of aerodynamics and

inertial forces have been carried out using the Brown University wind-tunnel facility, and a

force sensor1 positioned below the robot’s body. The experiments are focused on the motion

of the morphing-wings, i.e., wings extension and contraction by means of elbow’s motion. The

response of the morphing-controller is shown in Figure 7.2. As previously mentioned, two

scenarios are considered for testing the performance of the control architecture: i) non-flapping

with Vair = 0m/s), and ii) morphing+flapping at f = 2.5Hz with Vair = 5m/s. The term Vair

1Nano17 transducer ATI Industrial Automation, 0.318 gram-force of resolution, http://www.ati-ia.com/

156

Page 189: BaTboT: a biologically inspired flapping and morphing bat robot

7.2 Control performance

(a) (b)

(c) (d)

©Brown University©Brown University

©Brown University©Brown University

(e) (f)

Figure 7.2: Stills of morphing-wings’ control within the wind-tunnel. The wingbeat cycle is

composed by two phases: downstroke and upstroke. a) Beginning of the downstroke. The body of

the specimen is lined up in a straight line, elbow joint is∼ 58o, b) end of downstroke, the membrane

is cambered and the wings are still extended, elbow joint is ∼ 5o, c) middle of downstroke, the

wings are extended to increase lift, elbow joint is ∼ 20o, d) upstroke, the wings are folded to

reduce drag, elbow joint is ∼ 45o. Figures a-b illustrate the process to measure aerodynamics

loads using the force sensor located at the center of mass of the robot (below the body). Figures

c-d illustrate the process to measure inertial forces at the center of mass produced by both wings

(no aerodynamics loads caused by the membrane). e-f) show the beginning of downstroke and the

end of the upstroke without the membrane.

denotes the airspeed of the wind-tunnel. Performance is then evaluated in terms of: i) tracking

accuracy, ii) actuation speed, and iii) SMA fatigue issues.

157

Page 190: BaTboT: a biologically inspired flapping and morphing bat robot

7.2 Control performance

0 2 3 6

0.15

0.3

0.45

0.6

0.75

t [s]

I [A]

0 2 4 60

20

40

60

80

t [s]

3 [deg

]

0.1 0.2 0.3 0.40

25

50

75

t [s]

3 [deg

]

0.1 0.2 0.3 0.40

0.2

0.4

0.6

0.8

t [s]

Posit

ion er

ror

(Anti-slack)~10% of max. current

(Anti-overload)(a) (b)

(c) (d)

q3 at Vair = 0m/sq3,ref q3,ref

q3 at Vair = 0m/sq3 at Vair = 5m/s

error, Vair = 0m/serror,Vair = 5m/s

0 4

q3 [

deg]

q3 [

deg]

Isma

[A]

Figure 7.3: (Experimental) morphing-wings’ control response. a) Tracking of the elbow’s joint

trajectory at f = 2.5Hz, Vair = 0m/s i.e., no-wind. b) Close-up to a wingbeat cycle. The two plots

describe the control tracking regarding: i) Vair = 0m/s (same than plot-a), and ii) Vair = 5m/s.

c) Electrical current Isma delivered to the antagonistic SMA actuators, and regulated by the

anti-slack and anti-overload mechanisms. d) Position tracking errors from plot-b.

7.2.1.1 SMA accuracy and speed

Figure 7.3a-b shows the experimental results regarding the motion tracking of morphing-wing

trajectories at a wingbeat cycle of f = 2.5Hz. In order to analyze the accuracy of the controller,

7.3b details the time-evolution of q3 during a wingbeat cycle (t = 0.4s). In this figure, the

bio-inspired reference trajectory profile is denoted as q3,ref , in which the downstroke phase

takes 0.22s (wings extended), and the upstroke 0.18s (wings folded). Two experiments have

been carried out: i) morphing tracking with Vair = 0, and ii) with Vair = 5m/s. After

several trials, we observed that significant errors in position tracking mostly occurred during the

upstroke phase. The corresponding errors are shown in Figure 7.3d. During the upstroke, drag

forces caused high aerodynamics loads that introduced serious disturbances that are difficult to

completely reject.

One might expect that the Resistance-Motion relationship (see Algorithm 3) provides a

feasible indirect measurement of motion under any condition, but at higher speeds, variations in

SMA electrical resistance are presented during the entire wingbeat cycle. This fact is a problem

since the RM algorithm evaluates this resistance change only at the end of each wingbeat cycle,

158

Page 191: BaTboT: a biologically inspired flapping and morphing bat robot

7.2 Control performance

0

0.005

0.01

0.015

0.02

3 [Nm]

0

0.005

0.01

0.015

0

0.005

0.01

0.015

0

0.005

0.01

0 1 2 3 4

0.01

0.015

0.02

t [min]

3 [Nm]

0 4 8 12 160

0.004

0.008

0.012

0.016

t [min]

3 [Nm]

2.5Hz1.75Hz

1.35Hz1.1Hz

1.3Hz 1.23Hz

(a) (b)

(c)

0 5

0.002

0.004

0.006

0.008

0.005

0.01

0.015

0.02

Figure 7.4: (Experimental) Performance of the SMA actuator for longer periods of actuation:

a) Nominal operation at 1.3[Hz], b) Overloaded operation at 2.5[Hz], c) Output torque peaks

extracted from overloaded response in plot-b).

and this means that the slope correction of the RM function only takes place for the next cycle.

As a consequence, cumulative errors are introduced in the estimation of q3. These errors could

be reduced by introducing a prediction stage within Algorithm ??. In terms of actuation speed,

the implemented control architecture allowed the system to operate successfully at f = 2.5Hz.

7.2.1.2 SMA fatigue issues

Most of the previous experiments have been carried out for short periods of time. However,

contracting and extending the wings at 2.5Hz requires peaks of input power ∼ 3W , which

could cause the SMA to fatigue over time. Unfortunately for longer periods of SMA actuation,

fatigue issues are observed, causing the output torque of the SMA actuators to decrease over

time. As consequence, SMA performance in terms of actuation speed also decreases quickly

as a function of time. Figure 7.4a-b show the measured output torque curve produced by the

SMA actuators at both nominal and overloaded operation modes. The optimal performance

(f = 2 − 2.5[Hz]) can be maintained up to 1.5 minutes without compromising the actuation

speed (at minute 1.5, the reduction in output torque is about 4.5%). Figure 7.4c shows the

measured peaks of torque produced by the elbow joint during rotation. Each peak corresponds

to the four critical points highlighted in plot-b. Table 7.1 summarizes the data.

After 5 minutes of continuos overloaded operation, the wingbeat frequency has decreased

from f = 2.5Hz to 1.1Hz (stabilizing around 1Hz). This corresponds to ∼ 56% of performance

decrease in terms of actuation speed. On the other hand, -under nominal mode- (see Figure

159

Page 192: BaTboT: a biologically inspired flapping and morphing bat robot

7.2 Control performance

Table 7.1: Performance data of SMA actuation for longer periods of time.

Time Output torque Actuation speed % Performance-reduction1

Nominal initial 0.007[Nm] 1.3[Hz] –

16[min] 0.0068[Nm] 1.23[Hz] 5%

initial 0.018[Nm] 2.5[Hz] –

1.5[min] 0.015[Nm] 1.75[Hz] 30%

Overloaded 3[min] 0.011[Nm] 1.35[Hz] 46%

5[min] 0.008[Nm] 1.1[Hz] 56%

1 % of reduction in the actuation speed.

7.4a), results have shown that the SMA actuators tend to stabilize around 1 − 1.2Hz. These

results confirm that constant output torque in a Migamotor actuator can be only maintained

under nominal operation, at SMA contraction speeds of 300ms (96). Once overloaded, the

critical fatigue point is presented about 1.5 minutes of continuos operation, however, it has

been observed that once the SMA actuator is completely cooled, it is able to raise the maximum

operating frequency of 2.5Hz during another 1.5 minutes. Table 7.1 details the performance

data. In particular, methods for removing or reducing fatigue phenomenon must be analyzed.

One of these methods could be based on investigations related to high-frequency responses of

SMAs and the possibility of using high-bandwidth control systems as a possible approach of

eliminating limit cycles. As demonstrated by (9), high-bandwidth force control could be a

solution.

7.2.2 Attitude control

This section presents wind-tunnel experimental results regarding the performance of the pro-

posed attitude controller in terms of:

1. evaluating the accuracy of the backstepping+DAF controller for tracking pitch and roll

references under the presence of external disturbances caused by aerodynamic loads at

airspeeds up to 5ms−1,

2. demonstrating the assumption of incrementing net body forces (Fnet) thanks to the wing

modulation driven by the backstepping+DAF controller.

7.2.2.1 Forward flight

During forward flight, the bat-robot flaps the wings symmetrically at the desired wingbeat

frequency f . The bio-inspired angular trajectories of the wings are shown in figure 4.5. Wing

160

Page 193: BaTboT: a biologically inspired flapping and morphing bat robot

7.2 Control performance

Figure 7.5: Forward flight control. Backstepping+DAF attitude tracking at: a)-b) roll and pitch

tracking with Vair = 5ms−1, c)-d) roll and pitch tracking with Vair = 2ms−1.

modulation produces pitching and rolling torques (τφ, τθ) due to wing inertias cause angular

accelerations on the body. Using the inertial model of the robot, it is possible to determine

both roll and pitch motions that are produced by these angular accelerations. The experiment

carried out in figure 7.5 uses the calculation of roll and pitch motions as input references for

the attitude controller (φd, θd). The goal is to assess the accuracy of the backstepping+DAF

method during the tracking of φd and θd when subjected to aerodynamic loads. As shown

in figure 7.5a-b, the backstepping+DAF is able to maintain both φ, θ oscillating around the

defined set-point. The roll set-point is 0o whereas the pitch set-point is 5.5o. This configuration

allows for the generation of positive inertial thrust that would drive the robot forward. For

this scenario the wind-tunnel airspeed has been set to 5ms−1. Figures 7.5c-d follow the same

procedure with the difference that pitch set-point has been set to 20o and the airspeed has been

decreased to 2ms−1. The insets show the attitude tracking errors caused by aerodynamic loads

that depend on the airspeed.

7.2.2.2 Turning flight

In turning flight, the roll angle must be modified to allow the robot to turn right or left. Figure

7.6 shows experimental results on how the robot behaves during this process. Roll references are

161

Page 194: BaTboT: a biologically inspired flapping and morphing bat robot

7.2 Control performance

disturbance

disturbance

disturbance

disturbance

Figure 7.6: Turning flight control. Backstepping+DAF attitude tracking at: a)-b) roll and pitch

tracking with Vair = 5ms−1, c)-d) roll and pitch tracking with Vair = 2ms−1.

defined of the form φd = a+ bsin(2πft), where a is the roll set-point angle, b is the amplitude

of the oscillation, and f the desired frequency. On the other hand, pitch references have been

set to θd = 0o. In figure 7.6a, the roll reference φd = a + 0.25sin(2πft) determines that the

robot must turn from right to left by following the set-point command: a = 7o (0 < t ≤ 4),

and a = −2.5o (4 < t ≤ 7). Disturbances have been induced into the system aimed at testing

the reliability of the controller. These disturbances are caused by small loss of lift forces that

occur during the contraction process of the wing. This issue can be observed in Figs. 7.6a-c at

the switching point when t = 4s. The loss in lift forces accounted for about 4% due to a loss of

tension of the wing membrane. In this work the loss of tension has not been quantified. This

issue could be solved by improving the anisotropic property of the wing membrane material.

Disturbances of this kind are difficult to immediately reject at high airspeeds (5ms−1), however

the backstepping+DAF controller has shown accurate performance in attenuating the amplitude

of these oscillations at nominal airspeeds of 2ms−1 (cf. figure 7.6c-d). Without the DAF, the

oscillations caused by the disturbances would be higher than those observed here.

162

Page 195: BaTboT: a biologically inspired flapping and morphing bat robot

7.3 Aerodynamics experiments

0 5 10 150

0.5

1

1.5

Angle of attack [deg]

CL an

d CD

0 5 10 154

6

8

10

12

Angle of attack [deg]

L/D

0 5 10 15 204.995

5

5.005

5.01

Measurements

V air [m

/s]

0 5 10 1500.20.40.60.8

11.2

Angle of attack [deg]

L and

D Fo

rces [

N]

CL non-morphingCL

morphing

CD morphing

CD non-morphing

(a) (b)

(c) (d)

L

D

Figure 7.7: (Experimental) Aerodynamics measurements. a) Comparison between lift and drag

coefficients (CL, CD) with and without the motion of the morphing-wings (Vair = 5m/s, wingbeat

frequency of f = 2.5Hz). b) Lift-to-drag ratio (L/D) as a function of the angle of attack (Vair =

5m/s, f = 2.5Hz ). c) Wind-tunnel airspeed measurements (Vair). d) Lift (L) and drag (D)

forces corresponding to plot-a (with morphing).

7.3 Aerodynamics experiments

An initial aerodynamic analysis was previously introduced in Figure 4.10. The goal of the

experiment was to quantify lift and drag coefficient values to include them into the inertial

model as aerodynamic loads. In that experiment, the wings were fixed to the robot at mid-

downstroke, whereas the angle of attack was increased up to 25o. Contrary, this section shows

aerodynamic performance of the entire robot, performing both flapping and morphing-wings’

motions within the wind-tunnel facility. The experiments carried out in Figure 7.7 are focused

on analyzing the changes in lift and drag that are induced by the morphing wing.

Figure 7.7a compares the obtained lift and drag coefficients when the wings of the robot

are: i) -static-, i.e., non-morphing motions, and ii) describing the bio-inspired wing trajectories,

i.e., flapping+morphing motions synchronized at a wingbeat frequency of f = 2.5Hz. The

morphing trajectory described by the elbows’ joints was shown Figure 7.3b, whereas the flapping

trajectory is a simple sinusoidal signal with amplitude of ±60o with f = 2.5Hz. In terms of

lift production, note in Figure 7.7a how the lift coefficient (CL morphing) is increased by about

46% compared to the non-morphing profile. We have observed that the major contribution in

163

Page 196: BaTboT: a biologically inspired flapping and morphing bat robot

7.4 Efficient Flight

increasing lift forces was due to the flapping motion, accounting for about 28%. The morphing-

motion’s contribution accounted for about 18% thanks to wing area maximization during the

downstroke. On the other hand, the drag coefficient is reduced by about 40% by means of

folding the robot’s wings during the upstroke. The angle of attack corresponding to these

measurements is 9o.

Another important measurement corresponds to the lift-to-drag ratio L/D (Figure 7.7b )

which determines the most efficient angle of attack for sustained flight. This angle corresponds

to 9o (straight flight). At this angle, the corresponding lift and drag forces are shown in Figure

7.7d, being the lift force: 0.97N , about 26% more than the bat’s weight-force of 0.77N . Table

7.2 details these results.

Table 7.2: Lift and drag measurements for an angle of attack of 10o and Vair = 5m/s

Lift force L[N ] Drag force D[N ] CL CD

non-morphing 0.67 0.17 1.03 0.26

flapping+morphing (f = 2.5Hz) 0.97 0.099 1.5 0.152

7.4 Efficient Flight

Wing inertia information is contained in the DAF with the purpose of including desired roll

and pitch angular accelerations within the control law. DAF allows for the proper modulation

of wing kinematics, which impacts the generation of both inertial and aerodynamic forces.

Therefore, the following experiments are aimed at comparing the attitude response of the system

with and without the DAF function, showing the benefits of the proposed controller in terms

of disturbance rejection and net force production. Figure 7.8 quantifies the improvement in

forward flight by carrying out measurements of: i) attitude tracking (φ, θ), ii) wing modulation

(q3), and iii) net force production (Fnet). Note how the backstepping+DAF is able to reject

disturbances caused by increasing the wind-tunnel airspeed up to 5ms−1 (top plots). Like the

experiments in figure 7.5, accurate roll and pitch tracking is ensured during forward and turning

flight. Each controller (with and without DAF) produce a different pattern of wing modulation

(q3) (middle plots). With DAF the upstroke portion of the wingbeat cycle generates less drag

thanks to the fact that the elbow joint contracts sufficiently to reduce the wing area at minimum

span. This clearly affects the value of Fnet and the inertial thrust components (fxb).Finally,

note in the bottom plots how with DAF the bias of net forces Fnet is about 23% higher thanks

to the proper modulation of the wing kinematics. Also, it is confirmed that inertial thrust

164

Page 197: BaTboT: a biologically inspired flapping and morphing bat robot

7.4 Efficient Flight

1 2 3t (s)

00

t (s)

t (s) t (s)

t (s) t (s)(a) (b)

dowstroke upstrokedowstroke upstroke

dowstroke upstroke

Figure 7.8: (Forward flight) benefits of the DAF to the proper modulation of wing-morphology

aimed at incrementing net forces (f = 2Hz, φref = 0o, θref = 10o): a) without the DAF, b)

with the DAF. Top: attitude tracking error and disturbance rejection; middle: detailed wing

modulation (elbow joint q3); bottom: net forces generated.

Table 7.3: List of parameters used for experiments in figure 7.8.

Backstepping AoA Vair f1 FL FD fzb2 fxb Fnet bias

with DAF 9o 5ms−1 1.5− 2.5Hz ¯0.97N ¯0.099N 0.77N [−4,+9]mN 0.11N

without DAF 9o 5ms−1 1.5− 2.5Hz ¯0.97N ¯0.12N 0.77N [−4.5,+7.5]mN 0.09N

1 applies for that range of wingbeat frequencies.2 Mb = 79g, no-battery included.

is positive during the upstroke and negative during the downstroke, causing the net force to

increase or decrease as a function of the wingstroke motion (see inset in figure 7.8b). Table 7.3

summarises the numerical data.

In the experiments of Figure 7.8, efficient flight was evaluated by comparing the difference

between wing modulation profiles (q3) and their impact into the generation of net forces (Fnet).

As mentioned, the differences were produced by comparing the response of the backstepping

attitude control with and without the incorporation of DAF terms into the control law. However

165

Page 198: BaTboT: a biologically inspired flapping and morphing bat robot

7.4 Efficient Flight

xo[m] yo[m]

Zo[m

]

(a)

(b)

(c)

(d)

(e)

q3 no-DAF-1 q3 no-DAF-2

q4 no-DAF-1 q4 no-DAF-2

0.022

0.0215

0.021

0.0205

0.02

Figure 7.9: Effects of different wing modulation profiles (no-DAF terms presented) on lift and

drag production (f = 2.5Hz, Vair = 5ms−1): a) Cartesian trajectory of the wingtip generated

during a wingbeat cycle measured with respect to the base frame {0}, b) wing modulation profile of

elbow joint q3 for different backstepping parameter values of λ2 and λ4 (no-DAF-1: λ2 = λ4 = 0.1,

no-DAF-2: λ2 = λ4 = 0.05), c) wing modulation profile of wrist joint q4 corresponding to the

same backstepping parameter values configuration from plot-b, d) (left) net forces and (right)

lift and drag coefficients generated with the backstepping parameter configuration no-DAF-1, e)

left) net forces and (right) lift and drag coefficients generated with the backstepping parameter

configuration no-DAF-2.

in both cases, the backstepping parameters shown in Table 6.2 were used (λ1, C1, λ2). The goal

of the following set of experiments in Figure 7.9 consists on observing the effects of backstepping

parameter tuning for the generation of wing modulation commands and also show how the

performance of flight is drastically decreased when DAF terms are not included and also the

166

Page 199: BaTboT: a biologically inspired flapping and morphing bat robot

7.4 Efficient Flight

Table 7.4: Summary of performance of backstepping+DAF control and its influence into wing

modulation.

Backstepping AoA Vair f1 FL FD fzb2 Fnet bias λ2, λ4

with DAF3 9o 5ms−1 1.5− 2.5Hz ¯0.97N ¯0.099N 0.77N 0.11N 0.02

without DAF4 9o 5ms−1 1.5− 2.5Hz ¯0.97N ¯0.12N 0.77N 0.09N 0.02

no-DAF-25 9o 5ms−1 1.5− 2.5Hz ¯0.71N ¯0.17N 0.77N 0.085N 0.05

no-DAF-16 9o 5ms−1 1.5− 2.5Hz ¯0.31N ¯0.12N 0.77N 0.021N 0.1

1 applies for that range of wingbeat frequencies.2 Mb = 79g, no-battery included.3,4 data from Figures 7.8 and 7.7.5,6 data from Figure 7.9.

original backstepping parameters are not suitable tuned.

In relation to the backstepping parameter tuning, it has turned out that the parameter

λ2 (roll control) and λ4 (pitch control); cf. Eq. (6.29) is key for the proper tuning of back-

stepping control response. Both parameters directly affect the error e2, having an impact into

the regulation of the joint speed. Whether λ2 and λ4 increase, the slope of the joint angular

function also increases, which consequently enable faster downstroke cycles. This can be ob-

served in the experiments from Figure 7.8 (medium plots) note both q3 profiles tend to have

similar downstroke and upstrokes periods and even with the DAF terms included (Figure 7.8b-

medium) downstroke phases take longer than the upstroke. This response is induced aimed at

incrementing lift forces when wings are extended during the downstroke and reduce drag when

wing are folded during the upstroke. In the practise, backstepping+DAF parameters has been

tuned as λ2 = λ4 = 0.02 aimed at inducing the aforementioned response. Contrary, in Figures

7.9b-c the parameters λ2 and λ4 have been increased: no-DAF-2 refers to λ2 = λ4 = 0.05 and

no-DAF-1 refers to λ2 = λ4 = 0.1. One can note that larger values for λ2 and λ4 induce faster

downstroke phases. Because wing modulation drastically affects both inertial and aerodynamic

responses, note in Figures 7.9d-e how net forces (Fnet) and lift coefficient (CL) profiles dras-

tically decreased compared to the experiments in Figure 7.8a-b (lower plots). The decrease

in performance is even worst because drag coefficients (CD) actually increase a bit, causing

aerodynamic frictions to increase. In conclusion, efficient flight can be achieved when original

backstepping control law is augmented with DAF terms (backstepping+DAF) and both λ2 and

λ4 parameters are lower. The influence of wing-inertia control into lift and net force production

can be really appreciated in Table 7.4 (this table completes the numerical values from 7.3 in

more detail).

167

Page 200: BaTboT: a biologically inspired flapping and morphing bat robot

7.5 Discussion of results: Towards efficient flight

7.5 Discussion of results: Towards efficient flight

7.5.1 Morphing-wing modulation

Taking inspiration from nature, and in particular the morphology and flight kinematics of

bats, we have proposed a biomechanical and control system design that takes advantage of the

morphing-wing capability of the bat flight apparatus. Although a conventional servo system is

used for the primary flapping mechanism, the proposed control enables the bat robot to perform

bio-inspired morphing-wing motions using Shape Memory Alloys used as artificial muscles for

wing retraction and extension. Experiments were carried out to analyze how to properly speed-

up the operation of the SMAs to ensure their feasible use for the application at hand.

In terms of control, the adapted anti-slack and anti-overload mechanisms proved to effec-

tively work in a position control scheme, by servo’ing SMA electrical resistance changes to

accurately estimate the morphing-motion of the wings. Thanks to the implemented Resistance-

Motion (RM) relationship, both mechanisms were analyzed and experimentally adjusted for

regulating the amount of input heating power to be delivered to the SMA artificial muscles.

The fact that our robot does not make use of any motion sensor to provide the angular po-

sition feed-back, and yet achieves satisfactory tracking errors (even in the presence of high

aerodynamics loads), represents a validation of this control approach.

7.5.2 Wing inertia for efficient flight

The results presented in this section demonstrate how the wings can considerably affect the

dynamics/aerodynamics of flight and how to take advantage of wing inertia information to

properly change wings’ geometry during flapping. This fact has been carefully modelled and

quantified for the prototype at hand. It has also been shown how an effective attitude control of

the bat-like robot can be achieved by changing wings’ kinematics in order to generate controlled

inertial forces. The robot’s body mass influence in the generation of pitching and rolling torques

has been quantified, and scaling factors that relate how both inertial quantities increase as

a function of the flapping frequency, were found. These factors allowed the tuning of the

backstepping+DAF control parameters, improving the attitude tracking against high external

disturbances produced by aerodynamic loads.

Moreover, the proposed control strategy was developed and tested for demonstrating the

assumption of incrementing net body forces thanks to the wing modulation driven by the

backstepping+DAF controller. Such increment is significant, about 23%. Part of that increment

168

Page 201: BaTboT: a biologically inspired flapping and morphing bat robot

7.5 Discussion of results: Towards efficient flight

0 1002.5

3

3.5

4

4.5

5

A x,b[m

/s2 ]

Percentage of wingstroke %

0 100Percentage of wingstroke %

A xbms

−2⎡ ⎣

⎤ ⎦

2.5

3

3.5

4

4.5

5

Figure 7.10: (Experimental) comparison of horizontal inertial acceleration (Axb) produced by

the wing modulation at f = 2.5Hz and Vair = 5ms−1: (above) biological data of C. brachyotis

specimen; several measurements reported in (24), (below) BaTboT; several estimations of Axb

based on measurements of inertial thrust fxb.

was provided by the increasing in lift and decreasing of drag forces, cf. Figure 7.7, however,

inertial acceleration or thrust also played and important role in effective net force production.

Figure 7.10 shows the results. For the specimen, cf. Figure 7.10-above, when comparing the

inertial acceleration throughout a wingbeat cycle, calculated as the difference between the

accelerations of the body markers and the centre of mass of the specimen, bats showed large

differences in horizontal peak inertial accelerations between low and high speed flights.

The results support the idea that inertial accelerations produced by the flapping motion of

relatively massive wings can considerably affect the kinematics of bat flight. Horizontal inertial

effects were maximal in both peak and mean accelerations at slow speeds for the downstroke as

well as for the upstroke phases of the wingbeat cycle and decreased as flight speed increased.

One possible explanation for the decrease in horizontal inertial acceleration with speed is that

the total horizontal excursion of the wing decreases with speed as a consequence of a more

vertical stroke plane angle observed at higher speeds. Thus, inertial effects can have important

implications on the way we interpret horizontal accelerations, particularly during slow flights,

cf. Figure 7.10-below.

169

Page 202: BaTboT: a biologically inspired flapping and morphing bat robot

8

Conclusions and Future Work

”Bat flight is fascinating, the engineering journey to accomplish it with BaTboT, even more”

8.1 General conclusions

BaTboT is the first prototype of its kind with the potential for achieving autonomous and sus-

tained flight. All the novel methodologies introduced in this thesis are aimed at achieving that

goal. Motivated by the potential behind bat flight and the lack of highly articulated morphing-

wing MAVs (not necessarily bat-like), BaTboT is definitively a step towards a new generation of

Micro Aerial Vehicles with tremendous dexterity and manoeuvrability that changing the wing’s

geometry enables. In pursuing this long-term vision, a hypothesis was declared:

Quantifying the effects of wing inertia in terms of thrust and lift production and

therefore including wing inertia information into the flight controller will allow

for the proper modulation of wing kinematics that finally would produce and

increase of net forces, thereby improving on flight efficiency.

To demonstrate and validate the aforementioned hypothesis, most of the methods for design,

modelling and control were based on an exhaustive and unprecedented analysis of its biological

counterpart, that finally provided the robust foundation to approach each state of BaTboT’s

development.

170

Page 203: BaTboT: a biologically inspired flapping and morphing bat robot

8.1 General conclusions

Design

A novel design framework relating kinematics/aerodynamics parameters with morphological

parameters was defined based on validated biological data, cf. Table 3.6. It shows how body

and wing mass influence on the proper definition of design criteria. Because BaTboT is the

first bat-like robot with highly articulated wings, this framework is key for future developments

of similar bat-like robots with different morphological parameters.

Modeling

Models for kinematics, dynamics, aerodynamics and actuation were defined and experimentally

validated. An inertial model allowed for the quantification of the influence of wing inertia into

robot’s maneuverability and the key role of proper wing modulation aimed at the production

of rolling and pitching torques for turning and forward flight.

Control

A Flight Control Architecture was defined. The proposed backstepping+DAF method for

attitude control has demonstrated to be key for achieving proper response in terms of attitude

stabilisation and tracking. More important, the assumption of incrementing net body forces

thanks to the wing modulation driven by the backstepping+DAF controller was proved. Such

increment was significant, about 23%. Also, aerodynamics were dramatically improved thanks

to the morphing-wing control, cf. Figure 7.7. In terms of lift production, the lift coefficient

was increased by about 46% thanks to the maximisation of the wing area during extension and

drag was reduced by about 40% by means of folding the robot’s wings during the upstroke.

Mimicking the way bats take advantage of inertial and aerodynamical forces produced by the

wings in order to both increase lift and maneuver is a promising way to design more efficient

flapping wings MAVs. This is a key factor for their effective use in practical applications where

the extremely low payload capacity limits the autonomy operation of flying machines in outdoor

scenarios.

The novel wing modulation strategy and attitude control methodology presented and vali-

dated in this thesis provide a totally new way of controlling flying robots that eliminates the need

of appendices such as flaps and rudders. These developments are a key step towards achieving

the first bat-like robot capable of sustained autonomous flight. The possibility of controlling the

171

Page 204: BaTboT: a biologically inspired flapping and morphing bat robot

8.2 Future Work

shape of the wings has great potential to improve the maneuverability of current Micro-Aerial

Vehicles. Also, the overall results highlight the importance of the incorporation of inertial effects

in future analyses of the kinematics of flapping locomotion and how inertial-based control for

attitude regulation is useful when having bio-inspired MAVs with an important wing-to-body

mass ratio.

8.2 Future Work

The need of new light and robust materials for robot fabrication is a top order for future

developments of BaTboT. Actually, ABS plastic has resulted too heavy for the application at

hand, therefore constraining the possibility of achieving sustained flight. Including electronics,

BaTboT has a mass of 125g with a wingspan of 53cm. The specimen, on the other hand, has a

mass of ∼ 80g, a proper tradeoff between mass and wingspan. Attempting to decrease BaTboT

mass will require the need of carbon fibre fabrication and the incorporation of light electronics

and sensors. Specially the LiPo battery that contributes about 36.8% of the overall mass and

only provides 800mAh of power, being inefficient for BaTboT requeriments. The development

of a low-mass high-power system based on super-capacitors, for instance, it is required.

The use of Shape Memory Alloys has been key in achieving light actuated morphing wings

but their power consumption and actuation speed are still a radical limitation. This thesis

explored how to speed-up SMA operation while maintaining the limits of power consumption,

however, future work dedicated to improve on SMA performance is required, specially in terms of

eliminating fatigue phenomenon by means of introducing high bandwidth controllers. Methods

for embedding force feedback into a single SMA actuator is a top order for future development

of these smart actuators.

To control wing modulation, further research should be directed to quantify the effects of the

incoming airflow on the wings, with the aim of adjusting wing morphology in a more efficient

way in order to dramatically reduce drag. Actually, the wings of biological bats have tiny hairs

that sense airflow conditions, and there is some evidence that this sensing apparatus in bats

contributes to their flight efficiency (8). Besides elbow’s contraction, bats’ 3-DoF wrist joint

also contributes in folding the digits towards the body. As a consequence, bats can reduce

their wingspan about 70% during the upstroke (28). In this work, we have attempted to mimic

part of that complexity, however, our robot is able to reduce its wingspan about 23% during

172

Page 205: BaTboT: a biologically inspired flapping and morphing bat robot

8.3 Thesis schedule

the upstroke. This mechanical limitation is due to the fact that the wrists’ joints are under-

actuated, contributing about 5% during wings’ contraction. Significant drag reduction still

remains a challenge. Also, further investigations regarding the highly-anisotropic property of

the wing membrane should be taken into account. Furthermore, future wing designs should

incorporate additional sensors, such as a flex sensor bound to the membrane or skeleton. This

will give real-time knowledge of the shape of the wing, allowing the wings’ shape in response

to external variables to be tailored.

8.3 Thesis schedule

The Gantt diagram details the tasks and milestones carried out during the thesis development.

2w1.1) Biomechanics: insects, birds, bats

6w1.2) Bat physiology

4w1.3) Bat aerodynamics

4w1.4) Smart materials

16w1) Nature's Flyers@UPM

3w2.1) Bat morphology

2w2.2) Bat muscles

2w2.3) Bat membranes

3w 1d2.4) Bat aerodynamics

3w 1d2.5) Bat skeleton structure

6w 1d2.6) Initial BAT-concept proposal

19w 3d2) Bat Kinematics of Flight@UPM

3) Milestone 1: Chapter 1-2-3 DocDocument Overview of tasks 1 and 2

2w 4d4.1) Choice of species: P. bachiotelis.

3w 2d 4h4.2) Morphological parameters

4w4.3) Wing kinematics of flight

2w4.4) Dynamics and aerodynamics

5w4.5) Proposed model for wing-muscle actuation

17w 1d 4h4) In-vivo Bat flight analysisIn-situ @Brown

5) Milestone 2: Chapter 4 DocDocument Overview of tasks 4

3w6.1) General morphology and kinematics

3w6.2) Dynamics Modeling

3w6.3) Aerodynamics Modeling

5w 1d6.4) SMA actuation modeling

3w 2d6.5) Open-loop testing

3w6.6) Model validation against in-vivo behavior

20w 3d6) BaTboT ModelingIn-situ @Brown

7) Milestone 3: Chapter 5 DocDocument Overview of tasks 6

8w 1d8.1) CAD development

5w 3d8.2) biomechanics insights

6w 1d8.3) ABS skeleton: body+wings

6w 4d8.4) The wing-membrane6w 4d8.5) Actuators and sensors

6w 4d8.6) Hardware onboard

40w 2d8) BaTboT design and biomechanics@UPM

9) Milestone 4: Chapter 6 DocDocument Overview of tasks 8

19w 2d10.1) Morphing wing control

13w 4d10.2) Flapping wing control

23w 4d10.3) Control architecture

23w 1d10.4) closed-loop testing.

80w 1d10) BaTboT Control@UPM

11) Milestone 5: Chapter 7 DocDocument Overview of tasks 8

38w 1d12.1) Performance analysis (wind-tunnel)

34w12.2) BaTboT model identification

20w 3d12.3) Investigation of 6DoF flight

29w12.4) Flight control and maneuvering.

121w 4d12) Experimental resultsIn-situ @Brown

13) Milestone 6: Final DocDocument Overview of the Project results

Title EffortNature's Flyers @UPM

Biomechanics: insects, birds, bats 100%

Bat physiology 100%

Bat aerodynamics 100%

Smart materials 100%

Bat Kinematics of Flight @UPM

Bat morphology 100%

Bat muscles 100%

Bat membranes 100%

Bat aerodynamics 100%

Bat skeleton structure 100%

Initial BAT-concept proposal 100%

Milestone 1: Chapter 1-2-3 Doc Document Overview of tasks 1 and 2

In-vivo Bat flight analysis In-situ @Brown

Choice of species: P. bachiotelis. 100%

Morphological parameters 100%

Wing kinematics of flight 100%

Dynamics and aerodynamics 100%

Proposed model for wing-muscle actuation 100%

Milestone 2: Chapter 4 Doc Document Overview of tasks 4

BaTboT Modeling In-situ @Brown

General morphology and kinematics 100%

Dynamics Modeling 100%

Aerodynamics Modeling 100%

SMA actuation modeling 100%

Open-loop testing 100%

Model validation against in-vivo behavior 100%

Milestone 3: Chapter 5 Doc Document Overview of tasks 6

BaTboT design and biomechanics @UPM

CAD development 100%

biomechanics insights 100%

ABS skeleton: body+wings 100%

The wing-membrane 100%

Actuators and sensors 100%

Hardware onboard 100%

Milestone 4: Chapter 6 Doc Document Overview of tasks 8

BaTboT Control @UPM

Morphing wing control 100%

Flapping wing control 100%

Control architecture 100%

closed-loop testing. 100%

Milestone 5: Chapter 7 Doc Document Overview of tasks 8

Experimental results In-situ @Brown

Performance analysis (wind-tunnel) 100%

BaTboT model identification 100%

Investigation of 6DoF flight 100%

Flight control and maneuvering. 100%

Milestone 6: Final Doc Document Overview of the Project results

Qtr 4 2009 Qtr 1 2010 Qtr 2 2010 Qtr 3 2010 Qtr 4 2010 Qtr 1 2011 Qtr 2 2011 Qtr 3 2011 Qtr 4 2011 Qtr 1 2012 Qtr 2 2012 Qtr 3 201

173

Page 206: BaTboT: a biologically inspired flapping and morphing bat robot

9

Publications

The development of this thesis has allowed the following scientific production, including JCR

referred journals, book chapters, conference proceedings and press articles.

9.1 Journals, book chapters and conference proceedings

Referred Journals (ISI-JCR)

1. (Q1) Colorado J., Barrientos A., Rossi C., Breuer, K, 2012. Biomechanics of smart wings in a bat robot:

morphing-wings using SMA actuators. Bioinspiration and Biomimetics. vol. 7, 036006 (16pp). (Topic

on BaTboT’s morphing-wing mechanism driven by SMA muscles)

2. (Q4) Colorado J., Barrientos A., Rossi C., 2011. (Bio-inspired Robots with Smart Muscles: Modeling,

Control, and Actuation) Msculos Inteligentes en Robots Biolgicamente Inspirados: Modelado, Control y

Actuacin. Revista Iberoamericana de Automtica e Informtica Industrial, vol. 8, No. 4, 385-396, published

by Elsevier. (Topic on SMA phenomenological modelling and force control)

3. (Q1) Colorado J., Barrientos A., Rossi C., 2012. Inertial Attitude Control of a Bat-like Morphing-wing

Micro Air Vehicle. Bioinspiration and Biomimetics. Accepted. (Topic on flight control using the

backstepping+DAF attitude approach)

4. (Q1) Colorado J., Barrientos A., Rossi C., 2012. Learning from bats: the influence of wing modulation

for efficient robotic flying vehicles. Journal of Royal Society Interface. Accepted. (Topic on kinematics

of bat flight applied to efficient flight)

5. (Q1) Rossi C, Colorado J., Coral W., Barrientos A., 2011. Bending Continuous Structures with SMAs:

a Novel Robotic Fish Design. Bioinspiration and Biomimetics, vol. 6, No. 4, 15pp. (Topic on SMA

identification and resistance control)

6. (Q1) Barrientos A., Colorado J., del-Cerro J., Martinez A., Rossi C., Sanz D., Valente J. 2011. Aerial

Remote Sensing in Agriculture: A practical approach to area coverage and path planning for fleets of mini

aerial robot. Journal of Field Robotics, vol 28, issue 5, pp 667-689. (Topic on backstepping+DAF

attitude control applied to conventional UAV)

Book chapters

1. Coral W., Rossi C., Colorado J., Barrientos, A., 2012. Muscle-like actuation in biologically inspired

robots: The state of arts review on SMA technology, in book Smart actuation and sensing systems -

174

Page 207: BaTboT: a biologically inspired flapping and morphing bat robot

9.2 Press and media

Recent advances and future challenges. In-Tech, ISBN: 978-953-307-990-4. (Literature review on

SMA technology used as actuators in bio-inspired robots

Conferences

1. Colorado J, Rossi C, Coral W, Barrientos A. 2012. SMA-driven modulation of highly articulated wings:

insights on efficient flight inspired by bats. IEEE International Conference on Intelligent Robots and

Systems -IROS2012. Workshop on Smart materials and alternative technologies for bio-inspired robots

and systems, Algarve, Portugal, Oct.7-12.

2. Colorado J., Barrientos A., Rossi C., 2011. Biomechanics of morphing wings in a Bat-robot actuated by

SMA muscles. Proceedings of the International Workshop on bio-inspired robots, Nantes, France, April

6-8.

3. Colorado J. Aerial, Terrestrial and Aquatic Robotics: Navigation and Control Issues. Invited talk at

Pontificia Universidad Javeriana, Cali, Colombia. August 2010.

9.2 Press and media

1. Physics Society USA: http://phys.org/news/2012-06-batbot-imitating-smart-maneuvers.html

2. The Verge USA: http://www.theverge.com/2012/6/4/3060768/batbot-flying-robot-drone

3. ABC.es: http://www.abc.es/20120609/ciencia/abci-crean-robot-vuela-como-201206091312.html

4. RTVE: http://www.rtve.es/noticias/20120130/robot-murcielago-musculos-artificiales/493938.shtml

175

Page 208: BaTboT: a biologically inspired flapping and morphing bat robot

References

[1] Bullen R.D. and McKenzie N.L. Scaling bat wingbeat fre-

quency and amplitude. Journal of Experimental Biology 205,

26152626, 2002. viii, 13, 16, 18

[2] R. Dudley. The biomechanics of insect flight: form, function, evolu-

tion. Princeton University Press, 2000. viii, 16, 18

[3] A.P. Willmott C.P. Ellington and A.L.R. Thomas. Flow visual-

ization and unsteady aerodynamics in the flight of the

hawkmoth. Transactions of the Royal Society of London Series B

Biological Sciences, 352(1351):303-316, 1997. viii, 16, 18

[4] Hill J. J. Smith. Bats: A Natural History. Austin: University of

Texas Press, 1984. viii, 20

[5] A. Bergou S. Swartz K. Breuer G. Taubin. Development of-

bat flight: morphologic and molecular evolution of bat

wing digits. In IEEE symposium on biological data visualization,

Providence, RI, USA, page 10, 2011. ix, 23

[6] T. Hubel N. Hristov S. Swartz K.S. Breuer. Time-resolved

wake structure and kinematics of bat flight. Exp Fluids,

46:933943, 2009. ix, 24

[7] R. Albertani T. Hubel S.M. Swartz K.S. Breuer J. Evers. In-Flight

Wing-Membrane Strain Measurements on Bats. In EX-

PERIMENTAL AND APPLIED MECHANICS, Conference Proceed-

ings of the Society for Experimental Mechanics Series Volume 17.

Springer., pages 437–445, 2011. ix, 24, 25

[8] S. Sterbing-DAngelo M. Chadha C. Chiu B. Falk W. Xian J. Barcelo

J. M. Zook and C. F. Moss. Bat wing sensors support flight

control. In Proceedings of the National Academy of Sciences, vol.

108, no. 27, pages 11291–11296, 2011. ix, 2, 24, 25, 172

[9] T. Yee. Fast Accurate Force and Position Control of Shape Memory

Alloy Actuators. PhD thesis, ANU College of Engineering and

Computer Science, 2008. ix, 11, 27, 28, 29, 30, 31, 122, 124,

153, 160

[10] G. Bunget. A Bio-Inspired Micro-Aerial Vehicle for Flapping Flight.

PhD thesis, North Carolina State University, 2010. ix, 2, 32,

33

[11] R.E. Callahan and E. Garcia. Bio-Inspired Bat Wing De-

sign and Fabrication. In Cornell University, Ithaca, NY, 14850,

page 10, 2010. ix, 3, 33

[12] N. Perez-Arancibia K. Ma K. Galloway J. Greenberg and R. Wood.

First controlled vertical flight of a biologically inspired

microrobot. Bioinsp. Biomim. 6, 036009 (11pp), 2011. ix, 34

[13] Wildschek A. Grunewald M. Maier R. Steigenberger J. Judas M. Deli-

giannidis N. and Aversa N. Multifunctional morphing trailing

edge device for control of an all-composite, all-electric

flying wing aircraft. In 26th Congr. Int. Council of the Aero-

nautical Sciences (ICAS), page 6, 2008. ix, 34, 35

[14] Wilkie W. Bryant G. and High J. Low-cost piezocomposite ac-

tuator for structural control applications. In SPIE 7th

Annu. Int. Symp. on Smart Structures and Materials, page 7, 2000.

ix, 35, 36

[15] FESTO Company. http://www.festo.com/. ix, 35, 36

[16] Center for Objective Microelectronics and University of Michigan.

Biomimetic Advanced Technology. http://www.michigancmes.org/.

ix, 36

[17] Abdulrahim M. and Lind R. Flight testing and response char-

acteristics of a variable gull-wing morphing aircraft.

In AIAA Guidance Navigation and Control Conference and Exhibit,

page 7, 2004. ix, 1, 36

[18] Bowman J. Sanders B. Cannon B. Kudva J. Joshi S. and Weisshaar

T. Development of next generation morphing aircraft

structures. In 48th AIAA Structures, Structural Dynamics, and

Materials Conference, page 6, 2007. ix, 36, 37

[19] Extraterrestrial Flight. In: Low Re Aerodynamics on Aircraft In-

cluding Applications in Emerging UAV Technology. RTO-AVT von

Karman Institute for Fluid Dynamics Lecture Series Nov. 24 - 28. ix,

36

[20] D.K Riskin J. Iriarte-Diaz K.M Middleton K.S Breuer and S.M Swartz.

The effect of body size on the wing movements of

pteropodid bats, with insights into thrust and lift pro-

duction. The Journal of Experimental Biology, vol. 213 (23) pp.

4110-4122, 2010. x, xi, xv, xxi, 1, 4, 8, 24, 39, 40, 41, 42, 43,

44, 47, 50, 54, 56, 109, 110

[21] D.K Riskin D.J Willis J. Iriarte-Diaz T.L Hedrick M. Kostandov J.

Chen D.H Laidlaw K.S Breuer and S.M SWARTZ. Quantifying the

complexity of bat wing kinematics. Journal of Theoretical

Biology, vol. 254, no. 3, pp. 604615, 2008. x, 1, 44

[22] Watts P. Mitchell E.J and Swartz S.M. A computational model

for estimating the mechanics of horizontal flapping

flight in bats: model description and validation. J. Exp.

Biol, 204, 2873-2898, 2001. xi, 8, 39, 45, 50, 54, 55, 56, 58, 95

[23] Tian X. Iriarte-Diaz J. Middleton K. Galvao R. Israeli E. Roemer A.

Sullivan A. Song A. Swartz S. and Breuer K. Direct measure-

ments of the kinematics and dynamics of bat flight.

Bioinspir. Biomim. 1, S10-S18, 2006. xi, 4, 8, 17, 39, 50, 54, 56

[24] J. Iriarte-Diaz D.K. Riskin D.J Willis K.S Breuer and S.M Swartz.

Whole-body kinematics of a fruit bat reveal the influ-

ence of wing inertia on body accelerations. Journal of

Experimental Biology, vol. 214, no. 9, pp. 1546-1553, 2011. xi, xii,

xx, 1, 4, 8, 13, 17, 39, 50, 54, 56, 66, 68, 84, 137, 169

[25] Arnold J. Song Daniel K. Riskin Sharon M. Swartz Rye M. Waldman

and Kenneth S. Breuer. Aerodynamic Behavior of Compli-

ant Membranes as Related to Bat Flight. In In Proceed-

ings of American Institute of Aeronautics and Astronautics, page 13,

2010. xv, 110

[26] A.J. Bergou S. Swartz K.S. Breuer G. Taubin. 3D Reconstruc-

tion and Analysis of Bat Flight Maneuvers from Sparse

Multiple View Video. In In Proceedings of the 1st IEEE Sympo-

sium on Biological Data Visualization, page 4, 2011. xvii, 4, 137,

138

[27] Swartz S.M. Groves M.D. Kim H.D. and Walsh W.R. Mechanical

properties of bat wing membrane skin. J. Zool. vol. 239,

pp. 357-378, 1996. 1, 22

[28] Swartz S.M. Bishop K.L. and Ismael-Aguirre M. Dynamic complexity

of wing form in bats: implications for flight performance. In Func-

tional and Evolutionary Ecology of Bats (ed. Z. Akbar, G.

McCracken and T. H. Kunz), pp. 110- 130. Oxford: Oxford

University Press, 2005. 1, 2, 172

[29] Norberg L U.M. Brooke A.P. Trewhella W.J. Soaring and non-

soaring bats of the family Pteropodidae (fying foxes,

Pteropus spp.): wing morphology and flight perfor-

mance. J. Exp. Biol. 203 (3), 651-664, 2000. 1

[30] Gomez J.C. and Garcia E. Morphing unmanned aerial vehi-

cles. Smart Mater. Struct, vol. 20, no. 10, pp 1-16, 2011. 1

[31] Wickenheiser A. and Ephrahim G. Optimization of Perching

Maneuvers Through Vehicle Morphing. Journal of Guid-

ance, Control, and Dynamics, vol. 31, no. 4, 2008, pp. 815-823.,

2008. 1

176

Page 209: BaTboT: a biologically inspired flapping and morphing bat robot

REFERENCES

[32] Chakravarthy A. Grant D. and Lind R. Time-Varying Dynamics

of a Micro Air Vehicle with Variable-Sweep Morphing.

Journal of Guidance, Control, and Dynamics, Vol. 35, No. 3,, pp.

890-903, 2012. 1

[33] Bilgen O. Marqui C. Kochersberger K. and Inman D. Macro-fiber

composite actuators for flow control of a variable cam-

ber airfoil. J. Intell. Mater. Syst. Struct, vol. 22, pp. 81-91, 2011.

1

[34] Khan M. Strelec J. Lagoudas D. and Yen J. Design and imple-

mentation of a shape memory alloy actuated reconfig-

urable airfoil. J. Intell. Mater. Syst. Struct, vol. 14, pp. 257-73,

2003. 1

[35] Popov A. Grigorie L. and Botez R. Closed-Loop Control Vali-

dation of a Morphing Wing Using Wind Tunnel Tests.

J. Aircr, vol. 47, pp.1309-1317, 2010. 1

[36] Vos R. DeBreuker R. Barrett R. and Tiso P. Morphing wing

flight control via postbuckled precompressed piezoelec-

tric actuators. J. Aircr, Vol. 44, pp.1060-9, 2007. 1

[37] Rakotomamonjy T. Ouladsine M. Le Moing T. Longitudinal mod-

elling and control of a flapping-wing micro aerial vehi-

cle. Control Engineering Practice, Vol. 18, pp.679-690, 2010. 1

[38] Doman D.B. Michael W. Oppenheimer and Sigthorsson D.O. Wing-

beat Shape Modulation for Flapping-Wing Micro-Air-

Vehicle Control During Hover. Journal of Guidance, Control,

and Dynamics, vol. 33, no. 3, pp. 724-739, 2010. 1

[39] Doman D.B. Michael W. Oppenheimer and Sigthorsson D.O. Nature

inspired flight-beyond the leap. Bioinsp. Biomim, vol. 5, no.

4, p040201, 2010. 1

[40] Lentink D. Muller U.K. Stamhuis E.J. Kat R. Gestel W. Veldhuis

L.L.M. Henningsson P. Hedenstrom A. Videler J.J. and Leeuwen J.L.

How swifts control their glide performance with mor-

phing wings. Nature Letters, vol. 446, pp.1082-1085, 2007. 1

[41] Swartz S.M. Galvao R. Iriarte-Diaz J. Israeli E. Middleton K. Roemer

R. Tian X. and Breuer K. Unique characteristics of aerody-

namics of bat flight evidence from direct visualization

of patterns of airflow in the wakes of naturally flying

bats. Integ. Comp. Biol, vol. 45, pp.1080, 2005. 1

[42] Hedenstrom A. Johansson L.C. and Spedding G.R. Bird or bat:

comparing airframe design and flight performance.

Bioinsp. Biomim, vol. 4, no.1, pp. 5001, 2009. 2

[43] Pons J.L. Emerging Actuator Technologies. John Wiley and Son,

2007. 2, 13, 25, 69

[44] P. Ouyang R. Clement W.J. Zhang and G.S. Yang. Micro Motion

Devices Technology: the State of Arts Review. Journal

of Advanced Manufacturing Technology Vol. 38, no. 56, pp. 463478,

2007. 2, 69

[45] V. Bundhoo E. Haslam B. Birch and E. J. Park. A shape mem-

ory alloy-based tendon-driven actuation system for

biomimetic artificial fingers, part I: design and eval-

uation. Robotica, vol. 27, no. 1, p. 131, 2008. 2, 69

[46] Rossi C. Colorado J. Coral W. Barrientos A. Bending Continu-

ous Structures with SMAs: a Novel Robotic Fish De-

sign. Bioinspir. Biomim. Special issue: biomimetics of movement,

Vol. 6, No. 4, 2011. 2, 69

[47] S. Yang and S. Seelecke. FE analysis of SMA-based bio-

inspired bonejoint system. Smart Mater. Struct., vol. 18, no.

10, p. 104020, 2009. 2, 32, 33

[48] E. A. Leylek J. E. Manzo and E. Garcia. Bat-Inspired Wing

Aerodynamics and Optimization. J. Aircraft, vol. 47, no. 1,

pp. 323328, 2010. 3, 33

[49] Thollesson M. and Norberg U.M. Moments of inertia of bat

wings and body. Journal of Experimental Biology, vol. 158, 1991,

pp. 19-35, 1991. 5, 138

[50] Alexander D.E. Nature’s Flyers: Birds, Insects, and the Biomechanics

of Flight. The Johns Hopkins University Press, 2002. 8, 13

[51] Swartz S.M. Skin and bones: the mechanical properties of bat wing

tissues; in Bats: Phylogeny, Morphology, Echolocation, and Conser-

vation Biology. in Kunz, T.H., P.A. Racey (eds), Smithsonian

Institution Press, 1998. 8, 13

[52] Swartz S.M. Bennett M.B. Carrier D.R. Wing bone stresses in

free flying bats and the evolution of skeletal design for

flight. Nature 359, 726729., 1992. 8, 20

[53] Denavit J. and Hartenberg R.S. A Kinematic Notation for

Lower Pair Mechanisms Based on Matrices. Trans. ASME

J. Applied Mechanics, vol. 22, pp. 215221, 1955. 9, 51, 52

[54] Phillips W.F. Mechanics of Flight. Hoboken, NJ: Wiley, 2004. 9,

51

[55] R. Featherstone. Rigid Body Dynamics Algorithms. Springer, New

York, 2008. 9, 48, 54, 61

[56] M.H. Elahinia. Effect of System Dynamics on Shape Memory Alloy Be-

havior and Control. PhD thesis, Virginia Polytechnic Institute

and State University, 2004. 9, 28, 49, 71, 73

[57] Neuweiler G. The Biology of Bats. Oxford University Press, 2000.

13

[58] Vaughan T.A. Flight Patterns and Aerodynamics. Biology of Bats.

Academic Press, 1970. 13, 22

[59] Winter Y. and O. Von Helversen. The energy cost of flight: do

small bats fly more cheaply than birds? J Comp Physiol

B 168: 105-111, 1998. 13

[60] Norberg. U. Energetics of Flight. Avian Energetics and Nutri-

tional Ecology 199-249, 1996. 13

[61] Norberg. U. Aerodynamics, kinematics, and energetics of

horizontal flapping flight in the longeared bat Plecotus

auritus. J. exp. Biol. 65, 179-212, 1976. 13

[62] M. Kohl. Shape Memory Microactuators. Springer, 2004. 13, 25,

30

[63] Pennycuick C.J. Bird Flight Performance - A Practical Calculation

Manual. Oxford University Press, 1989. 17, 20

[64] Thomas S.P. and Suthers R.A. The physiology and energetics

of bat flight. J. Exp. Biol. 57, 317335., 1972. 19

[65] Norberg U.M. 3D Reconstruction and Analysis of Bat Flight Maneu-

vers from Sparse Multiple View Video. Berlin, Springer-Verlag,

1990. 20

[66] Sears K.E. R.R. Behringer J.J. Rasweiler L.A. Niswander. SDevel-

opment ofbat flight: morphologic and molecular evo-

lution of bat wing digits. In Proc Nat Acad Sci USA, pages

6581–6586, 2006. 23

[67] Y.H. Teh and R. Featherstone. An Architecture for Fast and

Accurate Control of Shape Memory Alloy Actuators.

The International Journal of Robotics Research, vol. 27 no. 5, 595-

611, 2008. 27, 29, 30, 122, 135, 153

[68] V. Brailovski S. Prokoshkin P. Terriault and F. Trochu. Shape mem-

ory Alloys: Fundamentals, Modeling and Applications. Universite

du Quebec, Ecole de technologie superieure, 2003. 28

[69] I. Mihalcz. Fundamental characteristics and design

method for nickel-titanium shape memory alloy. Pe-

riodica Polytechnica Ser. Mech. Eng. 45(1):75-86, 2001. 28

[70] M.H. Elahinia and M. Ahmadian. An enhanced SMA phe-

nomenological model: I. The shortcomings of the exist-

ing models. Smart Mater. Struct. 14, 1297-1308, pp. 1-13, 2005.

28, 29, 49, 71, 73

[71] Tanaka. A thermomechanical sketch of shape memory

effect: one-dimensional tensile behavior. Res. Mech., Int.

J. Struct. Mach. Mater. Sci. 18 251-63, 1986. 28, 70, 72

177

Page 210: BaTboT: a biologically inspired flapping and morphing bat robot

REFERENCES

[72] Katsutoshi Kuribayashi. A new actuator of a joint mecha-

nism using Ti-Ni alloy wire. International Journal of Robotics

Research, 4(4):4758, 1986. 28, 29

[73] Katsutoshi Kuribayashi. Improvement of the response of an

SMA actuator using a temperature sensor. The Interna-

tional Journal of Robotics Research, vol. 10, no. 1, pp.1320, 1991.

28, 29, 70

[74] S. Hirose K. Ikuta and Y. Umetani. A new design method of servo-

actuators based on the shape memory effect. Theory and Practice

of Robots and Manipulators, MIT Press, 1985. 28

[75] K. Ikuta M. Tsukamoto and S. Hirose. Mathematical model and

experimental verification of shape memory alloy for de-

signing micro actuator. In Micro Electro Mechanical Systems,

page 103108, 1991. 28

[76] D. Madill and D. Wang. Modelling and L2-stability of a

shape memory alloy position control system. In IEEE

Transactions on Control Systems Technology, page 473481, 1998.

29

[77] C. Liang and C. A. Rogers. Design of shape memory alloy

springs with applications in vibration control. Journal of

Vibration and Acoustics, 115:129 135, 1993. 29

[78] Brinson L.C. One-dimensional constitutive behavior of

shape memory alloys: thermomechanical derivation

with non-constant material functions and redefined

martensite internal variable. J. Intell. Mater. Syst. Struct.

4 22942., 1993. 29, 71

[79] D. Grant. Accurate and Rapid Control of Shape Memory Alloy Actu-

ator. PhD thesis, McGill University, 1999. 29, 30

[80] D. Honma Y. Miwa N. Iguchi. In Micro robots and micro

mechanisms using shape memory alloy. In In Integrated

Micro Motion Systems, Micro-machining, Control and Applications,

Nissan-Toyota Conference, page 6, 1989. 29

[81] B. Gorbet and R. Russell. A novel differential shape mem-

ory alloy actuator for position control. Robotica, 13:423430,

1995. 29

[82] Ikuta k. Tsukamoto M. and Hirose S. Shape memory alloy servo

actuator system with electric resistance feedback and

application for active endoscope. In IEEE International Con-

ference on Robotics and Automation, ICRA, page 427430, 1988. 29,

70

[83] D. Reynaerts and H. Van Brussel. Design aspects of shape

memory actuators. Mechatronics, 8:635656, 1998. 29

[84] N. Troisfontaine P. Bidaud and P. Dario. Control experiments on two

SMA based micro-actuators. Experimental Robotics, springer,

1998. 29

[85] Y. Sugiyama and S. Hirai. Crawling and jumping by a de-

formable robot. International Journal of Robotics Research, 25(5-

6):603620, 2006. 29

[86] Grant D. and Hayward V. Variable Structure Control of

Shape Memory Alloy Actuators. IEEE Systems and Con-

trol Magazine, vol. 9, no. 3, pp. 8088, 1997. 30

[87] Zhong Z.W. and Yeong C.K. Development of a Gripper us-

ing SMA Wire. Sensors and Actuators A: Physical, vol. 126, pp.

375-38, 2006. 30

[88] N.Ma G.Song and H. Lee. Position control of shape memory

alloy actuators with internal electrical resistance feed-

back using neural networks. Smart Mater. Struct. 13, 777-783,

2004. 30

[89] N.T. Tai and K.K. Ahn. Adaptive proportionalinte-

gralderivative tuning sliding mode control for a shape

memory alloy actuator. Smart Mater. Struct., vol. 20, no. 5, p.

055010, 2011. 30

[90] Jan Van Humbeeck. Non-medical applications of shape

memory alloys. Materials Science and Engineering A, 273-

275:134138, 1999. 31

[91] Siciliano B. and Khatib O. (eds). Springer Handbook of Robotics.

Springer, Berlin, 2008. 54

[92] J. Iriarte-Diaz and S.M SWARTZ. Kinematics of slow turn ma-

neuvering in the fruit bat Cynopterus brachyotis. The

Journal of Experimental Biology 211, 3478-3489, 2008. 58

[93] Rossi C. Coral W. Colorado J. Barrientos A. A Motor-less and

Gear-less Bio-mimetic Robotic Fish Design. In IEEE

International Conference on Robotics and Automation-ICRA, pages

3646–3651, 2011. 69

[94] Colorado J. Barrientos A. Rossi C. Biomechanics design of

smart wings in a Bat robot: morphing wings SMA con-

trol. Bioinspir. Biomim., 2012. 69

[95] E. Esfahani and M. Elahinia. Stable Walking Pattern for an

SMA-Actuated Biped. Mechatronics, IEEE/ASME Transac-

tions on, vol. 12, no. 5, pp. 534-541, 2007. 71

[96] Migamotors Company. http://www.migamotors.com/. 88, 89, 130,

160

[97] Colorado J. Barrientos A. Rossi C. (Bio-inspired Robots with

Smart Muscles: Modeling, Control, and Actuation)

Msculos Inteligentes en Robots Biolgicamente Inspira-

dos: Modelado, Control y Actuacin. Revista Iberoamericana

de Automtica e Informtica Industrial, Vol. 8, No. 4, 385-396, 2011.

147

[98] O. Harkegard. Flight Control Design Using Backstepping. PhD the-

sis, Linkopings universitet, Department of Electrical Engi-

neering, Sweden, 2001. 150

[99] S. Bouabdallah. Design and control of quadrotors with application to

autonomous flying. PhD thesis, Ecole Polytechnique Federale

de Lausanne (EPFL), Switzerland, 2007. 150

178

Page 211: BaTboT: a biologically inspired flapping and morphing bat robot

Declaration

I herewith declare that I have produced this thesis document without the prohibited assistance of

third parties and without making use of aids other than those specified; notions taken over directly

or indirectly from other sources have been identified as such.

The thesis work was conducted from 2010 to 2012 under the supervision of Prof. Antonio Barrientos

Cruz, PhD, and Prof. Claudio Rossi, PhD.

Madrid, Spain 2012

Page 212: BaTboT: a biologically inspired flapping and morphing bat robot

10

Annexes

10.1 Floating-base forward kinematics Matlab-code

1 function [Xb,Vb,Vdb] = fb Fkine(DH,q,qd,qdd)

2 %Add a virtual 6DOF kinematic joint, q, qd, add, contain the overall joint trajectory profile of the wing

3 fb q = [0 0 0 0 0 0 q]'; fb qd = [0 0 0 0 0 0 qd]'; fb qdd= [0 0 0 0 0 0 qdd]';

4 %obtain 3x3 rotation matrix that relates joint {1} with the floating base {0}5 c4 = cos(q(4)); s4 = sin(q(4)); c5 = cos(q(5)); s5 = sin(q(5)); c6 = cos(q(6)); s6 = sin(q(6));

6 r = [ c5*c6, c4*s6+s4*s5*c6, s4*s6−c4*s5*c6;7 −c5*s6, c4*c6−s4*s5*s6, s4*c6+c4*s5*s6;

8 s5, −s4*c5, c4*c5 ];

9 p = fb q(1:3);

10 p skew = [ 0, −p(3), p(2);

11 p(3), 0, −p(1);12 −p(2), p(1), 0 ];

13 %coordinate transform from fixed to floating base coordinates

14 Xb = [ r, zeros(3); −r*p skew, r ];

15 % Derivative of r

16 rd = [ 1 0 sin(fb q(5));

17 0 cos(fb q(4)) −sin(fb q(4))*cos(fb q(5));

18 0 sin(fb q(4)) cos(fb q(4))*cos(fb q(5)) ];

19 omega = rd*fb qd(4:6); pd = fb qd(1:3);

20 %spatial velocity of the floating base expressed in fixed−base coordinates.

21 Vb = [ omega; pd+cross(p,omega) ];

22 c4d = −sin(fb q(4))*fb qd(4); s4d = cos(fb q(4))*fb qd(4);

23 c5d = −sin(fb q(5))*fb qd(5); s5d = cos(fb q(5))*fb qd(5);

24 rdd = [ 0 0 s5d;

25 0 c4d −s4d*c5−s4*c5d;26 0 s4d c4d*c5+c4*c5d ];

27 omegad = rd*fb qdd(4:6) + rdd*fb qd (4:6); pdd = fb qdd(1:3);

28 %spatial acceleration of the floating base expressed in fixed−base coordinates.

29 Vdb = [ omegad; pdd+cross(pd,omega)+cross(p,omegad) ];

180

Page 213: BaTboT: a biologically inspired flapping and morphing bat robot

10.2 Floating-base inverse dynamics Matlab-code

10.2 Floating-base inverse dynamics Matlab-code

1 function [FT,Vdb] = IDf(Xb,Vb,Vdb,q,qd,qdd,Faero,flag)

2 % Evaluating for each wing

3 for k=1:2

4 %Floating base contribution

5 Vel(k).V(1)=Vbf;

6 Acel(k).A(1)=Abf+[0 0 0 0 0 −9.81]';7 H=[0;0;1;0;0;0];

8 %Forward recurrence: Spatial Velocity and Acceleration

9 for i=1:n

10 V(i)=Vel(k).V(i);

11 A(i)=Acel(k).A(i);

12 %Calculating 3x3 transformations

13 [r,p]=fkine2(q(i)); %computes r and p with basic forVrd kinematics using Hd−parameters14 %6x6 notation

15 psk=[0 −p(3),p(2);p(3),0,−p(1);−p(2),p(1),0];16 R=[r,zeros(3,3);zeros(3,3),r]; P=[eye(3),−psk;zeros(3,3),eye(3)];17 %Compute Spatial Velocity

18 Vel(k).V(:,i+1)=P*R'*V(:,i)+H*Qd(i);

19 %Spatial Acceleration terms

20 V skew1=[0 −V(3,i+1),V(2,i+1);V(3,i+1),0,−V(1,i+1);−V(2,i+1),V(1,i+1),0]; %actual

21 V skew1 6=[V skew1,zeros(3,3);zeros(3,3),V skew1];

22 Hd=V skew1 6*H;

23 V skew2=[0 −V(3,i),V(2,i);V(3,i),0,−V(1,i);−V(2,i),V(1,i),0];24 V skew2 6=[V skew2,zeros(3,3);zeros(3,3),V skew2];

25 Pd=V skew2 6*P';

26 %Compute Spatial Acceleration

27 Acel(k).A(:,i+1)=P*R'*A(:,i)+Pd*V(:,i)+Hd*qd(i)+H*qdd(i);

28 end

29 %Backward recurrence: spatial forces

30 if flag ==1

31 Force(k).F(n+1)=Faero; %include aerodynamics forces

32 else

33 Force(k).F(n+1)=zeros(6); %no−aero load (just accounting inertial effects

34 end

35 for i=n:−1:136 F(i)=Force(k).F(i);

37 [Jcm,m,s]=inertias(i); %computes the body tensor, mass, and distance to CM

38 s skew=[0 −s(3),s(2);s(3),0,−s(1);−s(2),s(1),0];39 %6x6 notation

40 S=[eye(3) s skew;zeros(3,3) eye(3)];

41 Sd=[zeros(3,3),V skew1 6*s skew;zeros(3,3),zeros(3,3)];

42 I=S'*[Jcm zeros(3,3);zeros(3,3) m(i)*eye(3)]*S;

43 id=V skew1 6*I;

44 %Compute Spatial forces

45 Force(k).F(:,i)=R*P'*F(:,i+1)+I*A(:,i+1)+(Id+I*Sd')*V(:,i+1);

46 end

47 end

48 R0 b =FB trans(q,Xb); %Computes transformation from frame {0} to body frame {b}49 FT=R0 b*Force(1).F(:)+R0 b*Force(2).F(:) %Floating base force FT

50 Ifb=I base() + Xb'*I*Xb; %Floating body inertia

51 Vdb=inv(Ifb)*FT; %Floating body acceleration

10.3 Floating-base forward dynamics Matlab-code

1 function [Abf] = FDf(FT,VdT,q,qd)

181

Page 214: BaTboT: a biologically inspired flapping and morphing bat robot

10.4 SMA phenomenological model Matlab-code

2 % Evaluating for each wing

3 for k=1:2

4 %Compute velocity dependent force terms C(q,qd)qd (using inverse dyn for this purpose)

5 [C tau] = IDf(zeros(6),zeros(6),zeros(6),q(k),qd(k),zeros(n),zeros(6));

6 % Step 1. Compute F = FT(q) − C(q,qd)qd

7 F(k).Ft = FT(k) − C tau;

8 %wing−recurrence9 for i=1:n

10 %6D transformations (R,P)

11 [r,p]=fkine2(q(i)); %computes r and p with basic forVrd kinematics using Hd−parameters12 %6x6 notation

13 psk=[0 −p(3),p(2);p(3),0,−p(1);−p(2),p(1),0];14 R=[r,zeros(3,3);zeros(3,3),r]; P=[eye(3),−psk;zeros(3,3),eye(3)];15 P = R*P';

16 %6D Inertia (I)

17 [Jcm,m,s]=inertias(i); %computes the body tensor, mass, and distance to CM

18 s skew=[0 −s(3),s(2);s(3),0,−s(1);−s(2),s(1),0];19 S=[eye(3) s skew;zeros(3,3) eye(3)];

20 Sd=[zeros(3,3),V skew1 6*s skew;zeros(3,3),zeros(3,3)];

21 I=S'*[Jcm zeros(3,3);zeros(3,3) m(i)*eye(3)]*S;

22 %Filling the high level spatial terms

23 if i<n

24 for a=0:5

25 for b=0:5

26 P h(((6*z)−5)+a,((6*g)+1)+b)=P(a+1,b+1); %P (6nx6x)

27 I h(c1+a,((6*c2)+1)+b)=I(a+1,b+1); %diag{I1,...In}(6nx6n)28 H h(u+a,i)=[0 0 1 0 0 0]'; %diag{H1,...Hn}(6nxn)29 end

30 end

31 z=z−1; g=g−1; u=u+6; c1=c1+6; c2=c2+1;

32 end

33 end

34 % Step 2. Compute the system overall mass M

35 M(k).MT = H h'*inv(P')*I h*inv(P)*H h;

36 % Step 3. Compute joints accelerations of the wings Qdd

37 Qdd(k).QT = inv(M(k).MT)*F(k).Ft;

38 end

39 % Step 4. Projecting the joint acceleration of both wings onto the bat's body

40 R2 1 =FB trans(q,Xbf); %Computes transformation from wing to body frame

41 H base = zeros(6*n,1);

42 H base(1:6) = ones(6,1);

43 Qdd T = H base*(Qdd(1).QT + R2 1*Qdd(2).QT);

44 % Step 5. Projecting spatial accelerations of both wings onto bat's body

45 Vd T = H base*(VdT(1) + R2 1*VdT(2));

46 % Solving the floating base spatial acceleration

47 Abf = Vd T−Qdd T;

10.4 SMA phenomenological model Matlab-code

1 function [T,strain,stress,theta] = SMA phenomenologicalModel(I,step,Time)

2 To = 20; %ambient temperature [C]

3 m = 0.00014; %SMA mass [Kg]

4 R = 8.5; %SMA initial resistance [Ohms]

5 Lo = 0.085; %link length

6 ro = 0.0025; %Link joint radius

7 % Fixed Parameters

8 Cp = 0.2; %Specific heat of wire

9 Ac = 0.0004712; %SMA wire?s circumferential area per unit length (150um)

10 hc = 150; %Heat convection coefficient

182

Page 215: BaTboT: a biologically inspired flapping and morphing bat robot

10.4 SMA phenomenological model Matlab-code

11 t = 0:step:Time; %Time vector

12 %Initial conditions

13 T(1) = To(1); %Initial Temperature [C]

14 stress(1) = 75; %Initial stress [MPa]

15 strain(1) = 0.04; %[MPa]

16 Text = To(1);

17 %**************************************************************************18 %Evolution during Heating

19 %Temperature [C]

20 p = length(t); cont2 = 1; cont = 1;

21 tempo =1;

22 for i=1:p−123 T(i+1) = step*((I*I*R)−hc*Ac*(T(i)−Text))+T(i); %Heating Temperature

24 cont = cont+1;

25 if tempo ≤ (length(To)−1)26 if cont > p/(length(To)−1);27 cont2 =cont2+1;

28 Text = To(cont2);

29 cont = 1;

30 end

31 tempo = tempo+1;

32 end

33 end

34 %Stress computing as a function of temperature

35 As = T(1); Af = T(i+1); aA = pi/(Af−As);36 bA = −aA/10.3; %10.3 is the effect stress constant on Austenite temperatures [MPa.1/C]

37 p = length(T);

38 for j=1:p−139 stress(j+1) = step*(((0.55+1120*(1/(Af−As)))*((T(j+1)−T(j))/step))/(1+1120*(1/(Af−As))))+stress(j);

%Computing stress [MPa]

40 end

41 %Martensite fraction computing and its derivative:

42 p = length(stress);

43 for k=1:p−144 M(k) = 0.5*(cos(aA*(T(k)−As)+bA*stress(k))+1); %Martensite fraction during heating

45 dM(k) = −0.5*(sin(aA*(T(k)−As)+bA*stress(k)))*(aA*((T(k+1)−T(k))/step)+bA*((stress(k+1)−stress(k))/step));46 T h(k) = T(k);

47 end

48 Austenite = M;

49 %Strain computing as a function of stress, temperature, and Marsenite fraction

50 p = length(M);

51 for u=1:p−152 strain(u+1) = (step/75000)*(((stress(u+1)−stress(u))/step)−0.55*((T(u+1)−T(u))/step)+1120*((M(u+1)−M(u))/step))+strain

%Computing strain [MPa]

53 end

54 %Kinematics model (SMA attached to a link)

55 p = length(strain); Theta(1) = 0; ΔY = Lo;

56 for w=1:p−157 Theta(w+1) = (−step*((Lo*((strain(w+1)−strain(w))/step))/(2*ro)))+Theta(w);58 end

59 Theta = Theta*(180/pi);

60 %**************************************************************************61 %Evolution during Cooling

62 %Temperature

63 i = i+1; i flag = i+1;

64 cont = 1; %flag counter used for knowing how many steps are required in cooling phase

65 Ms = T(i);

66 while (T(i) > (To+0.5))

67 T(i+1) = step*(−hc*Ac*(T(i)−To))+T(i); %Cooling temperature

68 t(i+1) = t(i)+step; %Filling time vector with the cooling phase

69 i = i+1; cont = cont+1;

183

Page 216: BaTboT: a biologically inspired flapping and morphing bat robot

10.5 Control code programmed into the Arduino

70 end

71 %Stress computing as a function of temperature

72 Mf = T(i); aM = pi/(Ms−Mf);73 bM = −aM/10.3; %10.3 is the effect stress constant on Austenite temperatures [MPa.1/C]

74 j = j+1;

75 j flag = j+1;

76 for j2=i flag:cont−177 stress(j+1) = step*(((0.55+1120*(1/(Ms−Mf)))*((T(j2+1)−T(j2))/step))/(1+1120*(1/(Ms−Mf))))+stress(j);

%Computing stress [MPa]

78 j = j+1;

79 end

80 %Martensite fraction computing and its derivative:

81 k = k+1; j2 = i flag; k flag = k; temp = 1;

82 for k2=j flag:cont−183 M(k) = 0.5*(cos(aM*(T(j2)−Mf)+bM*stress(k2)))+0.5; %Martensite fraction during heating

84 Martensite(temp) = M(k);

85 T c(temp) = T(j2);

86 j2 = j2+1; k = k+1; temp = temp+1;

87 end

88 %Strain computing as a function of stress, temperature, and Marsenite fraction

89 u = u+1; j2 = i flag; k2 = j flag;

90 for u2=k flag:(cont−3)91 strain(u+1) = (step/28000)*(((stress(k2+1)−stress(k2))/step)−0.55*((T(j2+1)−T(j2))/step)+1120*((M(u2+1)−M(u2))/step))+

%Computing strain [MPa]

92 j2 = j2+1; k2 = k2+1; u = u+1;

93 end

94 end

10.5 Control code programmed into the Arduino

1 //

2 // control.cpp

3 //

4 //

5 // Created by Julian Colorado on 06/04/11.

6 // Copyright 2011 MyCompanyName . All rights reserved.

7 //

8

9 //#include <iostream>

10 #include <Wire.h>

11 #include <Servo.h>

12

13 int gyroResult[3], accelResult[3]; //Store raw sensor output

14 int loopsPerServoUpdate = 2; //Update servos once every loopsPerServoUpdate loops

15 int loopCounter = 0; //Used for driving servos once every loopsPerServoUpdate loops

16 float timeStep = 0.01; //Main loop should run at 1/timeStep Hz

17 float biasGyroX, biasGyroY, biasGyroZ; //Bias values

18 float biasAccelX, biasAccelY, biasAccelZ; //for bias corrections

19 float gyroX, gyroY, gyroZ; //Values after bias

20 float accelX, accelY, accelZ; //correction in degrees

21 float pitchGyro = 0; //Pitch according to gyro

22 float pitchAccel = 0; //Pitch according to accelerometers

23 float pitchPrediction = 0; //Output of Kalman filter

24 float rollGyro = 0; //Roll according to gyro

25 float rollAccel = 0; //Roll according to accelerometers

26 float rollPrediction = 0; //Output of Kalman filter

27 float giroVar = 0.1; //Was 0.1

28 float ΔGiroVar = 0.1; //Was 0.1

29 float accelVar = 5; //Was 5

184

Page 217: BaTboT: a biologically inspired flapping and morphing bat robot

10.5 Control code programmed into the Arduino

30 float Pxx = 0.1; //Angle variance; was 0.1

31 float Pvv = 0.1; //Angle change rate variance; was 0.1

32 float Pxv = 0.1; //Angle and angle change rate covariance; was 0.1

33 float kx, kv; //Mysterious Kalman thingies

34 const int FIRArrayLength = 36; //Number of taps for low pass FIR filter

35 float accelXFIRArray[FIRArrayLength]; //Array for FIR calculations

36 float accelYFIRArray[FIRArrayLength]; //Array for FIR calculations

37 float accelZFIRArray[FIRArrayLength]; //Array for FIR calculations

38 float gyroXFIRArray[FIRArrayLength]; //Array for FIR calculations

39 float gyroYFIRArray[FIRArrayLength]; //Array for FIR calculations

40 float FIRFilter[FIRArrayLength] = { //FIR filter coefficients for: 0−4Hz: gain 5, ripple 12dB; 10−50Hz: gain 0.00000

41 0.00008103385604967716, 0.00027111799242077064, 0.0006854974917591413, 0.0014576515867607135,

42 0.0027530631363864716, 0.004753901748829375, 0.007635997857285217, 0.011541206789928664,

43 0.016544708776664904, 0.022626620444902264, 0.02964895502860427, 0.03734831257674501,

44 0.045342366754130024, 0.053157699234441635, 0.060270435821884946, 0.06616236694645264,

45 0.0703764023237606, 0.07257493449372505, 0.07257493449372505, 0.0703764023237606,

46 0.06616236694645264, 0.060270435821884946, 0.053157699234441635, 0.045342366754130024,

47 0.03734831257674501, 0.02964895502860427, 0.022626620444902264, 0.016544708776664904,

48 0.011541206789928664, 0.007635997857285217, 0.004753901748829375, 0.0027530631363864716,

49 0.0014576515867607135, 0.0006854974917591413, 0.00027111799242077064, 0.00008103385604967716

50 };51 const int zeros = 10;

52 const int poles = 10;

53 float accelXInput[zeros + 1], accelXOutput[poles + 1];

54 float accelYInput[zeros + 1], accelYOutput[poles + 1];

55 float accelZInput[zeros + 1], accelZOutput[poles + 1];

56 float gyroXInput[zeros + 1], gyroXOutput[poles + 1];

57 float gyroYInput[zeros + 1], gyroYOutput[poles + 1];

58 float minServoAngle = 50; //Limits to the throw

59 float maxServoAngle = 130; //of the servo arms

60 unsigned long timer; //Timer for calibrating loop time to timeStep seconds.

61 Servo leftServo, rightServo, throttleServo, rudderServo;

62 float pFactor = 1.5; //Multiplier between angle errors and servo signals

63 float leftCalc, rightCalc, leftServoSignal, rightServoSignal;

64

65 //Function for writing a byte to an I2C device

66 void writeTo(byte device, byte toAddress, byte val) {67 Wire.beginTransmission(device);

68 Wire.send(toAddress);

69 Wire.send(val);

70 Wire.endTransmission();

71 }72 //Function for reading num bytes from an I2C device

73 void readFrom(byte device, byte fromAddress, int num, byte result[]) {74 Wire.beginTransmission(device);

75 Wire.send(fromAddress);

76 Wire.endTransmission();

77 Wire.requestFrom((int)device, num);

78 int i = 0;

79 while(Wire.available()) {80 result[i] = Wire.receive();

81 i++;

82 }83 }84

85 //Function for reading gyroscopes

86 void getGyroscopeReadings(int gyroResult[]) {87 byte buffer[6];

88 readFrom(0x68,0x1D,6,buffer);

89 gyroResult[0] = (((int)buffer[0]) << 8 ) | buffer[1];

90 gyroResult[1] = (((int)buffer[2]) << 8 ) | buffer[3];

185

Page 218: BaTboT: a biologically inspired flapping and morphing bat robot

10.5 Control code programmed into the Arduino

91 gyroResult[2] = (((int)buffer[4]) << 8 ) | buffer[5];

92 }93

94 //Funcion for reading accelerometers

95 void getAccelerometerReadings(int accelResult[]) {96 byte buffer[6];

97 readFrom(0x53,0x32,6,buffer);

98 accelResult[0] = (((int)buffer[1]) << 8 ) | buffer[0];

99 accelResult[1] = (((int)buffer[3]) << 8 ) | buffer[2];

100 accelResult[2] = (((int)buffer[5]) << 8 ) | buffer[4];

101 }102

103 //FIR filter function

104 float FIR(float value, float array[]) {105 int i;

106 float FIROutput = 0;

107 for(i = FIRArrayLength−1; i > 0; i−−){108 array[i] = array[i−1];109 FIROutput += array[i] * FIRFilter[i];

110 }111 array[0] = value;

112 FIROutput += array[0] * FIRFilter[0];

113 return FIROutput;

114 }115

116 //IIR filter function (Low pass Bessel, 5Hz corner freq., 10th order, 100 Hz sample rate;

117 float IIR(float value, float xv[], float yv[]) {118 xv[0] = xv[1]; xv[1] = xv[2]; xv[2] = xv[3]; xv[3] = xv[4]; xv[4] = xv[5]; xv[5] = xv[6]; xv[6] = xv[7]; xv[7] = xv[8]

119 xv[10] = value / 5.681713320e+05;

120 yv[0] = yv[1]; yv[1] = yv[2]; yv[2] = yv[3]; yv[3] = yv[4]; yv[4] = yv[5]; yv[5] = yv[6]; yv[6] = yv[7]; yv[7] = yv[8]

121 yv[10] = (xv[0] + xv[10]) + 10 * (xv[1] + xv[9]) + 45 * (xv[2] + xv[8])

122 + 120 * (xv[3] + xv[7]) + 210 * (xv[4] + xv[6]) + 252 * xv[5]

123 + ( −0.0084477842 * yv[0]) + ( 0.1250487070 * yv[1])

124 + ( −0.8452516757 * yv[2]) + ( 3.4397583115 * yv[3])

125 + ( −9.3457158590 * yv[4]) + ( 17.7418428080 * yv[5])

126 + (−23.8769055880 * yv[6]) + ( 22.5422166190 * yv[7])

127 + (−14.3254341650 * yv[8]) + ( 5.5510863543 * yv[9]);

128 return yv[10];

129 }130

131 void setup() {132 //Variables used for removing zero bias

133 int totalGyroXValues = 0;

134 int totalGyroYValues = 0;

135 int totalGyroZValues = 0;

136 int totalAccelXValues = 0;

137 int totalAccelYValues = 0;

138 int totalAccelZValues = 0;

139 int i;

140

141 //LED on when calibrating

142 //pinMode(13, OUTPUT);

143 //digitalWrite(13, HIGH);

144

145 //Hook up the servos to digital pins 2 thru 5

146 leftServo.attach(5, 600, 2400);

147 rightServo.attach(4, 600, 2400);

148 throttleServo.attach(3, 600, 2400);

149 rudderServo.attach(2, 600, 2400);

150

151 //Configure pins for bat s throttle

186

Page 219: BaTboT: a biologically inspired flapping and morphing bat robot

10.5 Control code programmed into the Arduino

152 pinMode(A0, OUTPUT);

153 digitalWrite(A0, HIGH);

154 pinMode(A1, INPUT);

155 pinMode(A2, OUTPUT);

156 digitalWrite(A2, LOW);

157

158 //Set the servos

159 leftServo.write(90);

160 rightServo.write(90);

161 throttleServo.write(0);

162 rudderServo.write(90);

163 delay(50);

164

165 //Open I2C and serial communications

166 Wire.begin();

167 Serial.begin(115200);

168

169 //Apply settings to sensors

170 writeTo(0x53,0x31,0x09); //Set accelerometer to 11bit, +/−4g171 writeTo(0x53,0x2D,0x08); //Set accelerometer to measure mode

172 writeTo(0x68,0x16,0x1E); //Set gyro to +/−2000deg/sec, 1kHz internal sampling rate and 5Hz low pass filter

173 writeTo(0x68,0x15,0x0A); //Set gyro to 100Hz external sample rate

174

175 //Collect 50 measurements and calculate bias for sensors

176 delay(100);

177 for (i = 0; i < 50; i += 1) {178 getGyroscopeReadings(gyroResult);

179 getAccelerometerReadings(accelResult);

180 totalGyroXValues += gyroResult[0];

181 totalGyroYValues += gyroResult[1];

182 totalGyroZValues += gyroResult[2];

183 totalAccelXValues += accelResult[0];

184 totalAccelYValues += accelResult[1];

185 totalAccelZValues += accelResult[2];

186 delay(50);

187 }188 biasGyroX = totalGyroXValues / 50;

189 biasGyroY = totalGyroYValues / 50;

190 biasGyroZ = totalGyroZValues / 50;

191 biasAccelX = totalAccelXValues / 50;

192 biasAccelY = totalAccelYValues / 50;

193 biasAccelZ = (totalAccelZValues / 50) − 256;

194

195 //LED off when done calibrating

196 digitalWrite(13, LOW);

197 }198

199 void loop() {200 //Initialise the loop timer

201 timer = millis();

202

203 //Set the throttle to the bat

204 throttleServo.write(map(analogRead(1), 0, 1023, 0, 179));

205

206 //Get the raw sensor data

207 getGyroscopeReadings(gyroResult);

208 getAccelerometerReadings(accelResult);

209

210 //Remove bias from raw sensor data and convert to g and deg/s respectively.

211 accelX = (accelResult[0] − biasAccelX) / 256;

212 accelY = (accelResult[1] − biasAccelY) / 256;

187

Page 220: BaTboT: a biologically inspired flapping and morphing bat robot

10.5 Control code programmed into the Arduino

213 accelZ = (accelResult[2] − biasAccelZ) / 256;

214 gyroX = (gyroResult[0] − biasGyroX) / 14.375;

215 gyroY = (gyroResult[1] − biasGyroY) / 14.375;

216

217 //Try to remove the motor and rotor vibrations from the data with a low−pass FIR/IIR filter

218 accelX = IIR(accelX, accelXInput, accelXOutput);

219 accelY = IIR(accelY, accelYInput, accelYOutput);

220 accelZ = IIR(accelZ, accelZInput, accelZOutput);

221 gyroX = IIR(gyroX, gyroXInput, gyroXOutput);

222 gyroY = IIR(gyroY, gyroYInput, gyroYOutput);

223

224 //Calculate angles from the sensor data

225 pitchAccel = atan2(accelY, accelZ) * 360.0 / (2*PI);

226 pitchGyro = pitchGyro + (gyroX * timeStep);

227 pitchPrediction = pitchPrediction + (gyroX * timeStep);

228 rollAccel = atan2(accelX, accelZ) * 360.0 / (2*PI);

229 rollGyro = rollGyro − gyroY * timeStep;

230 rollPrediction = rollPrediction − (gyroY * timeStep);

231

232 //Do the Kalman filter

233 Pxx += timeStep * (2 * Pxv + timeStep * Pvv);

234 Pxv += timeStep * Pvv;

235 Pxx += timeStep * giroVar;

236 Pvv += timeStep * ΔGiroVar;

237 kx = Pxx * (1 / (Pxx + accelVar));

238 kv = Pxv * (1 / (Pxx + accelVar));

239 pitchPrediction += (pitchAccel − pitchPrediction) * kx;

240 rollPrediction += (rollAccel − rollPrediction) * kx;

241 Pxx *= (1 − kx);

242 Pxv *= (1 − kx);

243 Pvv −= kv * Pxv;

244

245 //Use the Kalman output in a controller, limiting the maximum servo arms throw

246 leftCalc = pitchPrediction − rollPrediction;

247 rightCalc = pitchPrediction + rollPrediction;

248 leftServoSignal = 90.0 − (pFactor * leftCalc);

249 rightServoSignal = 90.0 + (pFactor * rightCalc);

250 if (leftServoSignal > maxServoAngle) {leftServoSignal = maxServoAngle;};251 if (leftServoSignal < minServoAngle) {leftServoSignal = minServoAngle;};252 if (rightServoSignal > maxServoAngle) {rightServoSignal = maxServoAngle;};253 if (rightServoSignal < minServoAngle) {rightServoSignal = minServoAngle;};254

255

256 //Controller

257 float pFactor = 3;

258 float minServoAngle = 50;

259 float maxServoAngle = 130;

260 float pitch, roll, leftCalc, rightCalc, leftCalcFIR, rightCalcFIR, leftServoSignal, rightServoSignal;

261 pitch = pitchPrediction;

262 roll = rollPrediction;

263 leftCalc = pitch − roll;

264 rightCalc = pitch + roll;

265 leftCalcFIR = FIR(leftCalc, leftServoFIRArray);

266 rightCalcFIR = FIR(rightCalc, rightServoFIRArray);

267 leftServoSignal = 90.0 − (pFactor * leftCalcFIR);

268 rightServoSignal = 90.0 + (pFactor * rightCalcFIR);

269 if (leftServoSignal > maxServoAngle) {leftServoSignal = maxServoAngle;};270 if (leftServoSignal < minServoAngle) {leftServoSignal = minServoAngle;};271 if (rightServoSignal > maxServoAngle) {rightServoSignal = maxServoAngle;};272 if (rightServoSignal < minServoAngle) {rightServoSignal = minServoAngle;};273 leftServo.write((int)round(leftServoSignal));

188

Page 221: BaTboT: a biologically inspired flapping and morphing bat robot

10.6 Flight Control Matlab environment

274 rightServo.write((int)round(rightServoSignal));

275

276

277 //Update signal every loopsPerServoUpdate loops, aiming for once every 20ms

278 if (++loopCounter ≥ loopsPerServoUpdate) {279 leftServo.write((int)round(leftServoSignal));

280 rightServo.write((int)round(rightServoSignal));

281 loopCounter = 0;

282 }283

284 //Make the loop last timeStep seconds

285 timer = millis() − timer;

286 timer = (int)(round(timeStep * 1000)) − timer;

287 delay(timer);

288 }

10.6 Flight Control Matlab environment

189

Page 222: BaTboT: a biologically inspired flapping and morphing bat robot

10.6 Flight Control Matlab environmentW

ing

traje

ctor

ies G

ener

ator

Mod

els:

kin

emat

ics,

dyna

mic

s, m

echa

nics

Use

r ope

ratio

n

Out

er lo

op: A

ttitu

de c

ontro

l

Inne

r loo

p: m

orph

ing-

win

g co

ntro

l

Com

mun

icat

ion

data

I sma

2R s

ma

u_heating u_

heating

Figure

10.1:Closed-loopenvironmen

twithsliding-m

odemorphingcontrol.

Sourc

e:th

eauth

or.

190

Page 223: BaTboT: a biologically inspired flapping and morphing bat robot

10.6 Flight Control Matlab environment

Win

g tra

ject

orie

s Gen

erat

or

Mod

els:

kin

emat

ics,

dyna

mic

s, m

echa

nics

Use

r ope

ratio

n

Out

er lo

op: A

ttitu

de c

ontro

l

Inne

r loo

p: m

orph

ing-

win

g co

ntro

l

Com

mun

icat

ion

data

I sma

2R s

ma

u_heating u_

heating

Figure

10.2:Closed-loopenvironmen

twithPID

morphingcontrol.

Sourc

e:th

eauth

or.

191